Wiring the World: A History of the Concept in the US Earth Sciences, 1982-1989

by

Jenifer Patricia Barton

A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Graduate Department of the Institute for the History and Philosophy of Science and Technology University of Toronto

© Copyright by Jenifer Patricia Barton, 2020

Wiring the World: A History of the Earth System Concept in the US Earth Sciences, 1982-1989

Jenifer Patricia Barton

Doctor of Philosophy

Graduate Department of the Institute for the History and Philosophy of Science and Technology University of Toronto

2020 Abstract

Earth scientists today tend to view the planet as an integrated system comprised of interconnected components in the air, land, water, and biota. Earth scientists overwhelmingly use one particular phrase to describe this understanding of the planet: the “Earth system.” How did it become possible to conceive of the Earth as a system? When did “Earth system” become a common phrase? What were the principal factors that led to the concept’s later entrenchment?

This dissertation addresses these questions by examining the emergence of the “Earth system” concept among US scientists in the 1980s. While the “global” capacities afforded by post-World

War II Earth observing satellites and modeling may have been necessary for the conception of the Earth as a system, this dissertation argues that they were not sufficient. There is an important sociological component to the history of the Earth system.

The Earth system concept has bureaucratic origins that trace to the mid-1980s and the work of a small group of scientists, the Earth System Sciences Committee. This Committee— formed by NASA’s Advisory Council—developed and promoted a research program called

“Earth system science” that would take the whole planet as an object of study. Earth system science failed to gain extensive contemporary support, but the committee’s phrase “Earth system” was widely adopted. I argue that the “Earth system” phrase became entrenched despite

ii the failure of the larger project not because it was well defined but because it was vague. By the

1960s, scientists increasingly perceived satellites and global computer models as supporting the idea that the Earth had interconnected parts that required interdisciplinary study. There was, however, little agreement about how to express this imprecise idea. The phrase “Earth system” was vague enough to adequately fill this semantic void. It served as a boundary object between different scientific disciplines, with enough interpretive flexibility to be narrowly defined by specialists while at the same time being broad enough to facilitate interdisciplinary communication. Vagueness, not analytical precision, thus facilitated the early spread and widespread adoption of the Earth system concept and contributed to its later entrenchment in the

Earth sciences.

iii

Acknowledgments

I would first and foremost like to thank my supervisor—Chen-Pang Yeang—and my supervisory committee—Rebecca Woods, Edward Jones-Imhotep, and Matt Farish—for their support and advice. I owe a special thanks to Jay Foster for helping me work through many nebulous ideas and research threads.

This project would have looked quite different if it were not for the help of Laura Hoff, NCAR's archivist. Her answer to a simple email query completely transformed the trajectory of this research. I also benefited significantly from the assistance of Steve Garber, Colin Fries, and Elizabeth Suckow at NASA's Historical Reference Collection.

Thank you to the Social Sciences and Humanities Research Council of Canada for awarding me the Joseph-Armand Bombardier Canada Graduate Scholarship. This project was also supported by a Grant-in-Aid from the Friends of the Center for the History of Physics, American Institute of Physics. Greg Good was instrumental in securing this support. Greg Good, along with Jim Fleming and Ron Doel, provided invaluable intellectual help with this project.

Thank you to the IHPST graduate community. The department went through some ups and downs over the years, but my fellow graduate students were a steadfast source of companionship, relief, humour, and advice.

Lastly, I would like to thank my family for their unwavering support. This project would not have possible without them.

iv

Table of Contents

Abstract ii

Acknowledgements iv

List of Figures vi

List of Abbreviations vii

Introduction 1

Chapter One: The Emerging Threads: US Earth Sciences from the 1960s to the 1980s 34

Chapter Two: The and Death of NASA's Global Habitability Initiative 90

Chapter Three: The Earth System Sciences Committee: Constructing a Research Program 128

Chapter Four: The Earth System Sciences Committee: Promoting a Research Program 206

Chapter Five: The Strength of Vagueness: The AGU and the Spread of the Earth System 278

Epilogue: Mostly Harmless 341

Bibliography 355

v

List of Figures

Figure 0.1 - Google ngram plotting the comparative usage of “Earth system” in Google’s English corpus of books, 1900-2000, p. 3.

Figure 1.1 - The coupled --ice-earth climatic system. (US Committee for GARP, Understanding Climatic Change, 1975), p. 79.

Figure 2.1 - UNISPACE ‘82 stamp from the People’s Republic of China, 1982. (Reprinted with permission from PostBeeld), p. 121.

Figure 3.1 - Earth system wiring diagram from the ESSC’s Closer View (1988), depicting Earth processes occurring on timescales of decades to centuries. (Design by InterNetwork, Inc | Payson R. Stevens; reprinted with permission from NASA), p. 137.

Figure 3.2 - Earth system wiring diagram from the ESSC’s Closer View (1988), depicting Earth processes occurring on timescales of thousands to millions of years. (Design by InterNetwork, Inc | Payson R. Stevens; reprinted with permission from NASA), p. 191.

Figure 4.1 - NAS International Geophysical Year poster (1958) depicting scientific activities in the world’s . (Reprinted with permission from the National Academy of Sciences, courtesy of the National Academies Press, Washington, DC), p. 232.

Figure 4.2 - NASA’s from Space poster for “Phytoplankton Abundance” (1983) as designed by Payson Steven’s InterNetwork graphic design company. (Design by InterNetwork, Inc | Payson R. Stevens; reprinted with permission from NASA), p. 235.

Figure 4.3 - The ESSC’s Earth system science logo found on its reports and other media products. (Design by InterNetwork, Inc | Payson R. Stevens; reprinted with permission from NASA), p. 240.

Figure 4.4 - The “ESS Blue” serves as the background colour for all ESSC material, including its Closer View (1988) report. (Design by InterNetwork, Inc | Payson R. Stevens; reprinted with permission from NASA), p. 242.

Figure 4.5 - The simplified Earth system wiring diagram, for timescales of decades to centuries, from the ESSC’s Overview (1986) report. (Design by InterNetwork, Inc | Payson R. Stevens; reprinted with permission from NASA), p. 249.

Figure 4.6 - The Earth system science poster designed for the ESSC’s press conference, June 26, 1986. (Design by InterNetwork, Inc | Payson R. Stevens; reprinted with permission from Payson Stevens), p. 265.

vi

List of Abbreviations

AAAS American Association for the Advancement of Science AEC Atomic Energy Commission AGU American Geophysical Union AIP American Institute of Physics AMS American Meteorological Society AR6 IPCC Sixth Assessment Report CDR dioxide removal CESAS NAS’ Committee on and Applications from Space CES NAS’ Committee on Earth Sciences CFC Chlorofluorocarbon COPUOS United Nations Committee on the Peaceful Uses of Outer Space DDT Dichlorodiphenyltrichloroethane DOD Department of Defence DOE Department of Energy DST GARP Data Test ECD Electron capture detector ENIAC Electronic Numerical Integrator and Computer ENSO El Niño-Southern Oscillation EOS Earth Observing System EOSDIS Earth Observing System Data and ERBE Earth Radiation Budget Experiment ESA Ecological Society of America ESMWG Earth System Sciences Committee’s Earth Systems Modeling Working Group ESS Earth system Science ESSC Earth System Sciences Committee FAO Food and Agriculture Organization FGGE First GARP Global Experiment GARP Global Atmospheric Research Program GATE GARP Atlantic Tropical Experiment GCM General Circulation Model GGR Greenhouse gas removal GOES Geostationary Operational Environmental Satellites GPO Government Printing Office GREM Geopotential Research Explorer Mission HRC NASA Historical Reference Collection IAU International Astronomical Union ICSU International Council of Scientific Unions IEEE Institute of Electrical and Electronics Engineers IGBP International - Program IGY International Geophysical Year INI InterNetwork Inc. IPCC International Panel on Change ISLSCP International Satellite Land Surface Project JPL Jet Propulsion Laboratory NAC NASA’s Advisory Council vii

NAS National Academy of Sciences NASA National Aeronautic and Space Administration NCAR National Center for Atmospheric Research NIEO New International Economic Order NOAA National Oceanic and Atmospheric Administration NOSS National Oceanic Satellite System NRC National Research Council NSF National Science Foundation NWIO New World Information Order NWP Numerical Prediction OIES UCAR Office for Interdisciplinary Earth Studies OMB Office of Management and Budget OSSA NASA Office of Space Science and Applications OSTP Office of Science and Technology Policy OTA Office of Technology Assessment PEC AGU’s Planet Earth Committee PEI AGU’s Planet Earth Initiative PLATO Permanent Large Array of Terrestrial Observatories PMS Pantone Matching System SESAC NASA Advisory Council’s Space and Earth Science Advisory Committee SET Serial Endosymbiosis Theory SGC Systematics General Corporation SRM Solar radiation management SSB NAS’ Space Science Board TOPEX The Ocean Topography Experiment UARS Upper Atmosphere Research Satellite UCAR University Corporation for Atmospheric Research UNEP United Nations Environment Programme UNESCO United Nations Educational, Scientific, and Cultural Organization UNISPACE I First United Nations Conference on the Exploration and Peaceful Uses of Outer Space UNISPACE ‘82 Second United Nations Conference on the Exploration and Peaceful Uses of Outer Space USGCRP United States Research Program USGS United States Geological Survey WHOI Woods Hole Oceanographic Institute WCRP World Climate Research Program WMO World Meteorological Organization WOCE World Ocean Circulation Experiment WOW “Words of Wisdom” document

viii

Introduction

In March 2017, the US National Academy of Sciences (NAS) held its “Space Science Week,” an

annual opportunity for its numerous space science committees to meet and discuss issues and

new developments. Before the Committee on Earth Science and Applications from Space

(CESAS) convened for an open session, I spoke to a senior Earth science official at NASA and

said that I was examining the construction of the “Earth system” concept to learn how scientists

came to conceive of and study the Earth as a system. I wanted to understand the historical

trajectory of a phrase that is today nearly ubiquitous in the Earth and environmental sciences.1

The official immediately corrected me. He stated categorically that the Earth system has always

existed; scientists discovered the Earth system, they did not construct it.

Earth scientists use the “Earth system” to broadly describe the planet as a single,

interconnected system. The concept allows scientists to link together a range of seemingly

disparate phenomena into a more unified understanding of the planet. These linkages traverse

the traditional Earth science disciplines and studying them requires interdisciplinary

collaboration. For instance, phosphorus from algae dried at the southern edge of the Saharan

Desert, or from fires on the rest of the African continent, is carried by winds over 2,500

kilometers across the Atlantic Ocean and deposited in the Amazon rainforests, fertilizing the

vegetation with this locally-unobtainable nutrient.2 In the Pacific Ocean, the disappearance of the air pressure gradient between the east and west prevents the northward flow of cold waters

1 The “Earth system” phrase appears not just in scientific publications and presentations, but also in popular press coverage of and the Anthropocene, as well even in humanities and literature. 2 Recent research suggests that the source of this phosphorus is from African fires, from cooking or naturally occurring, rather than from dried algae from the Sahara. See: Anne E. Barkley, et al., “African Biomass Burning is a Substantial Source of Phosphorus Deposition to the Amazon, Tropical Atlantic Ocean, and Southern Ocean,” Proceedings of the National Academy of Sciences 116, no. 33 (13 Aug. 2019): 16216-21.

1 2

from the Humboldt Current, leading to a rise in ocean temperatures and a reduction in

precipitation for many parts of the Americas. These vast oceanic and atmospheric wobbles are

collectively known as the El Niño-Southern Oscillation (ENSO).3 Photoautotrophs in the oceans and on land transform water, incoming solar radiation, atmospheric carbon dioxide, and other nutrients into chemical energy via photosynthesis, thus transporting massive amounts of carbon from the atmosphere to the biosphere and .4 These are merely three accepted

examples of the many large-scale systemic interactions that occur on the planet. Earth scientists

today view the Earth as an integrated system, comprised of these and many other interactions

that link together the planet’s atmosphere, oceans, lands, ice, and biota.

The Earth system concept is now so familiar and so ubiquitous that for most Earth

scientists—including the NASA official I spoke with—it is undeniably the case that the Earth

system was always already “out there” in the world awaiting discovery. As a result, its conceptual and bureaucratic history is now almost completely forgotten or unheeded. However, the Earth system, like all concepts, has a history. The phrase “Earth system” appears to be ubiquitous in the Earth sciences today, but this has not always been the case. Prior to 1986, the

“Earth system” referred either to electrical-grounding schemes or to dry-earth composting toilets in vogue during the early twentieth century. It had nothing to do with the Earth and environmental sciences. Usage in Science and Nature, two of the most prominent non-specialist scientific journals, both demonstrate notable increases in the phrases’s use after 1985. Science has 906 hits post-1985 and only 23 hits for all of the preceding period, while Nature has 1,968

3 For instance, see: H. Chen, et al., “Enhancing the ENSO Predictability Beyond the Spring Barrier,” Scientific Reports 10, no. 984 (2020): 1-12. The pedosphere is the soil layer of the planet. 4 For instance, see: Danica L. Lombardozzi, et al., “Temperature Acclimation of Photosynthesis and Respiration: A Key Uncertainty in the ‐Climate ,” Geophysical Research Letters 42 (2015): 8624-31.

3 and 72 hits respectively.5 Google’s ngram viewer similarly reveals a surge in the relative frequency of “Earth system” since the mid-1980s in its English collection of books (Figure 0.1).6

Figure 0.1. Google ngram plotting the comparative usage of “Earth system” in Google’s English corpus of books, 1900-2000.

On or about 1986 the “Earth system” changed. The phrase came to refer to the planet as a system of interacting components.7 It was not an immediate and wholesale change, but I argue that this period marks the beginning of an underlying ontological shift. Through the promotional work of a NASA-funded group of US-based scientists attempting to develop a new Earth science research program, after 1986 the Earth became a new kind of thing. In the subsequent years, this conception of the planet gained such popularity that when scientists now use the “Earth system” no justification is required. It is used uncritically and unreflectively as a background concept.

For Earth scientists, the Earth is not studied as if it were a system. The Earth simply is a system.

5 These search results were obtained on 18 November 2019 from the Science at JStor (for 1880-2013) and the Science website (for 2014-present), and Nature’s online journal archive (1869-present). 6 Google’s ngram viewer searches for words or phrases within a “corpus of books” (e.g. “British English,” “English Fiction,” “French”) over a selected number of years and produces a graph which plots the trend in use of the particular word/phrase compared to the entirety of the corpus. See: “Google Books Ngram Viewer,” accessed 20 Dec 2019, https://books.google.com/ngrams. 7 Any pre-1986 exceptions using the “Earth system” phrase in its modern usage referred specifically to the work of the ESSC. For instance, see: National Research Council, An Implementation Plan for Priorities in Solar-system Space Physics (Washington, DC: National Academy Press, 1985), 35.

4

This dissertation presents a sociological and bureaucratic history of the Earth and

environmental sciences in the United States in the latter decades of the twentieth century to

understand how it became possible for scientists to conceptualize and study the Earth as a

system, one of the many ways that humans have “assembled” the planet as a whole.8 In particular, I focus on the crucial period from 1982 to 1989 when a nascent idea concretized into a specific phrase, the “Earth system.” To do this, I examine an important but previously overlooked component of the history of the Earth sciences. A clue comes from the unifying phrase itself. Earth scientists today do not describe the Earth as a system using a variety of different names. They collectively refer to the Earth as a system using a single common phrase.

The common phrase—Earth system—has a specific historical lineage. It has bureaucratic origins in a committee—the Earth System Sciences Committee (ESSC)—formed by the National

Aeronautics and Space Administration’s (NASA) Advisory Council in 1983 to develop a global

Earth science research program called Earth system science (ESS) that would use satellites and a systems approach to study the planet. The ESSC developed and promoted ESS in the mid- and

late-1980s in a variety of ways—including consultations with hundreds of scientists, the publication of glossy reports, a CNN-aired press conference, and even the printing of ESS t-

shirts—that created the conditions through which the Earth system concept gained traction and

entered common use.

This is not a story of scientists doing science at laboratory benches or in the field. This is the story of an exclusive group of US Earth scientists attempting to generate the scientific and political alignments that would enable a research program that took the whole planet as its

8 Rens van Munster and Casper Sylvest call these different holistic, universalizing, totalizing, and/or integrated conceptions of the whole planet as various assemblings of “globality.” These “globalities” can be, “multifarious, contradictory and deeply political.” See: Rens van Munster and Casper Sylvest, “Introduction,” in The Politics of Globality Since 1945: Assembling the Planet, eds. Rens van Munster and Casper Sylvest (New York: Routledge, 2016), 10.

5 research object. This is a story of how a core group of not just scientists but also graphic designers and administrators (along with their vast network of collaborators), working within the institutional mechanisms of NASA’s Advisory Council, developed an Earth system science research program. In the process, they contributed to the development and spread of the Earth system concept that instantiated the Earth as an interdisciplinary scientific object. It is a story full of disputes, frustrations, and tedium, but also much good humour, which often served as an outlet for the experienced hardships. The ESSC’s work provides an exemplar of post-World

War II interdisciplinary committee work. These committees not only played important practical roles in formulating research programs to attract government funding, but also provided avenues through which widely dispersed practitioners could come together to share ideas, formulate new ones, and then take those ideas back to their own constituencies. It provides an example of how, in the words of John Krige, scientific knowledge and ideas “move.”9 In the decade before the widespread adoption of the internet and electronic mail for communication, the role of committees like the ESSC in the development and spread of scientific ideas was invaluable.

In 1986, the “Earth system” concept began filling what I call, borrowing from Leo Marx, a “semantic void.”10 This “void” grew in US Earth science communities from the 1960s into the

1980s, produced in part by the new tools and techniques available in the latter part of the twentieth century and a lack of consensus regarding how to refer to an interconnected planet

(chapter one).11 Before the work of the ESSC, scientists referred to this growing conception of the planet in a variety of ways: the global system, the climatic system, the atmosphere-ocean-ice-

9 John Krige, ed., How Knowledge Moves: Writing the Transnational History of Science and Technology (: University of Chicago Press, 2019). 10 Leo Marx, “Technology: The Emergence of a Hazardous Concept,” Technology and Culture 51, no. 3 (Jul. 2010): 563. 11 Though this “void” presumably existed beyond US Earth science communities, this aspect falls beyond the scope of this dissertation.

6

earth climatic system, the terrestrial , the global ecosystem, the global environment,

the global biosphere, Spaceship Earth. If ’s Gaia hypothesis did not have

problematic connotations, “Gaia” might have filled the void (chapter one). If NASA’s Global

Habitability initiative had garnered widespread support in 1982 rather than rejection at

UNISPACE ‘82, it is possible that today “global ” would be in common use (chapter two).

However, it was ultimately a consequence of the extensive work of the ESSC that its Earth system phrase fully and finally filled the void. Although other phrases have not been entirely discarded, Earth and environmental scientists now refer to the Earth predominantly as the “Earth system.”12

The ESSC did not set out to provide common language to describe an interconnected

planet. Instead, the ESSC sought to generate support for a large-scale Earth science research

program called Earth system science (ESS). Despite its grand ambitions, ESS prioritized the

study of shorter term Earth processes like those studied by atmospheric scientists and

oceanographers, as well as space observations over in-situ measurements. This meant that many geologists and geophysicists—who studied longer term processes and were less reliant on space-

based platforms—found ESS as proposed problematic as it marginalized their interests and

needs, and could jeopardize their access to research resources. Despite such programmatic

disagreements, the ESSC’s “Earth system” phrase proved useful. It was vague enough to be

non-threatening to most disciplines and by the late 1980s readily entered widespread use. As a

result of its vagueness, the Earth system concept could serve as what Susan Leigh Star calls a

12 For example, there is recent research on the various histories of the “global environment,” though the phrase tends to be a historian, rather than actor, category. For a recent example, see: Studies in History and Philosophy of Science Part A, Special Issue: Experiencing the Environment, 70 (Aug. 2018): 1-86. There were only 388 hits for “global environment” in Science post-1985 (compared to the 906 hits for “Earth system”) and 625 hits in Nature for the same period (compared with 1,968 hits for “Earth system”). Further, many of the articles that use the phrase “global environment” are not research articles but intended for broader audiences.

7

“boundary object.” A boundary object has two distinct features. It is something that specialists

use in narrowly prescribed ways in specific contexts but also has enough interpretive flexibility

to transcend disciplinary boundaries and promote broad interdisciplinary communication despite

a lack of consensus about specifics.13 Practitioners from different Earth and environmental sciences used the “Earth system” to broadly describe an interconnected planet amenable to interdisciplinary research, while differently defining the particulars of the Earth system in specialist situations. Vague concepts might be philosophically undesirable but sometimes they prove practically useful. In some situations, a concept may be embraced not despite its vagueness but precisely because of it.

A HISTORICAL ONTOLOGY OF THE EARTH SYSTEM

This dissertation focuses on the emergence of the conditions for the possibility of thinking of and studying the Earth as a system. It examines how a small group of NASA-affiliated scientists developed and disseminated the nascent Earth system concept between 1982 and 1989. The

Earth system is not as exotic or amusing as the categories in Jorge Luis Borges’ fanciful

“Celestial Emporium of Benevolent Knowledge” and it is arguably less problematic than, say, certain psychiatric categories in the DSM-V that label humans “normal” and “pathological.”14

Yet, the Earth system concept is now so familiar and so ubiquitous that to many Earth scientists

it is an unquestionable fact that the Earth system was always “out there” in the world awaiting

discovery. Ola Uhrqvist has examined the process by which the Earth system concept and Earth

13 Susan Leigh Star, “The Structure of Ill-Structured Solutions: Boundary Objects and Heterogeneous Distributed Problem Solving,” in Readings in Distributed Artificial Intelligence, eds. Alan H. Bond and Les Gasser (San Mateo, CA: Morgan Kaufmann Publishers, 1988), 37-54. 14 Jorge Luis Borges, “The Analytical Language of John Wilkins,” in Other Inquisitions, 1937-1952, trans. Ruth L.C. Sims (Austin, TX: University of Texas Press, 2000), 103; American Psychiatric Association, Diagnostic and Statistical Manual of Mental Disorders: DSM-5 (Arlington, VA: American Psychiatric Association, 2013).

8 system science served as the scientific basis for identifying global environmental problems that became the focus of governments and other regulatory bodies in the 1990s and 2000s. Uhrqvist studies this transformation of the Earth system from an object of scientific knowledge into the political arena as an “object of concern” but Uhrqvist stops short of interrogating how the Earth system became an object of scientific inquiry in Earth science communities in the first place.15

Something cannot be a problem to be “solved” by governments before it is constituted as an object of knowledge, a process Michel Foucault calls “problematization.”16 Using never-before- analyzed archival documents from the ESSC, this dissertation fills the lacuna by examining how the Earth system first became a broad organizing concept in the Earth and environmental sciences.

The Earth system concept was not always already “out there” in the world awaiting discovery. It is not a natural or eternal category uncovered by human ingenuity and technique.

This is not to suggest that the Earth system is illegitimate or requires “unmasking.” As Donna

Haraway says, “To be ‘made’ is not to be ‘made up.’”17 Concepts and categories, like everything else, have histories. They emerge out of a variety of conditions: specific encounters in the world, material conditions, human concerns, technological capabilities, experimental and analytical techniques, disciplinary framings, interpersonal connections, and institutional interests.

Inspecting the history of concepts makes their local roots visible. It de-naturalizes concepts that are treated as obvious, inevitable, or never-changing. It brings the background to the fore. It

15 Ola Uhrqvist, “Seeing and Knowing the Earth as a System: An Effective History of Global Environmental Change Research as Scientific and Political Practice” (PhD diss., Linköping University, 2014), 5. 16 With the study of “problematization,” Foucault wants to know, “how and why certain things (behavior, phenomena, processes) became a problem” that could be known and therefore capable of eliciting attention and concern. See: Michel Foucault, “Discourse and Truth: The Problematization of Parrhesia (Six Lectures Given by Michel Foucault at the University of California at Berkeley, Oct-Nov. 1983),” accessed 10 Dec 2019, https://www.cscd.osaka-u.ac.jp/user/rosaldo/On_Parrehesia_by_Foucault_1983.pdf. 17 Donna J. Haraway, Modest_Witness@Second_Millennium.FemaleMan©_Meets_OncoMouseTM: Feminism and Technosciences, second edition (New York: Routledge, 2018), 99.

9

shows their contingent character. Since all concepts have unique histories, studying these

histories can reveal the specific conditions for the possibility of their emergence, the reasons for

how and why they emerged when they did. As Bruno Latour observes, technological artifacts

and scientific concepts are often analysed using very different standards. Technological artifacts

are never separated from the local, historical circumstances of their production. After an artifact

is newly created, we would never project its existence to all times, past and future. Yet, we often

do precisely this with scientific concepts by severing the connection between the concept and the

“local, material, and practical networks” of its production.18

Ian Hacking calls the study of the shifting ontological realities that come with the

emergence and disappearance of concepts “historical ontology.”19 Hacking readily admits that

this phrase, “is too self-important by half.” As “ontology” suggests, Hacking is interested in

“what there is” in the world: “Not just things, but whatever we individuate and allow ourselves

to talk about. That includes not only ‘material’ objects but also classes, kinds of people, and,

indeed, ideas.” Put simply, historical ontology focuses on any and all kinds of objects and

categories, “and what makes it possible for them to come into being.”20 Hacking traces the

phrase “historical ontology” to Michel Foucault’s essay “What is Enlightenment?” (1984) where

Foucault referred to the “historical ontology of ourselves.” Foucault suggested that it could form

the name for the study of how humans constitute themselves as objects of knowledge, subjects

acting on others, and as moral agents.21 Building on this suggestion (which Foucault himself did

18 Latour makes this call in a number of places, including his examination of how it is possible to meaningfully state that Ramses II, who died in 1213 BCE, died from a bacterium (Koch's bacillus) not identified by humans until 1882. See: Bruno Latour, “On the Partial Existence of Existing and Non existing Objects,” in Biographies of Scientific Objects, ed. Lorraine Daston (Chicago: University of Chicago Press, 2000), 247-69. 19 Ian Hacking, Historical Ontology (Cambridge, MA: Press, 2002). 20 Ibid, 1. 21 Michel Foucault, “What is Enlightenment?” in The Foucault Reader, ed. Paul Rabinow (New York: Pantheon Books, 1984), 32-50; Hacking, Historical Ontology, 2.

10

not further pursue), Hacking calls for the study of how all kinds of things are constituted and

how these new categories provide new ways for humans to “be” in the world. They provide new

horizons of possibilities within which human choice and action can occur. Historical ontology

is, for Hacking, the study of the ways in which these new spaces of possibility emerge as new

categories into which humans might be classified, or into which humans might choose to classify

themselves. Hacking refers to this process as the “looping effect” since human behaviour can

change based on a specific classificatory scheme in a way that a maple tree will not change its

behaviour based on any new categories that humans might devise for it.22 Hacking’s research

focuses primarily on the kinds of categories that might “make up people,” be they the categories

developed for early statistical studies by governments (for instance, statistics of deviancy, madness, and suicide) to psychiatric labels like “multiple personality disorder.”23 Once these

categories emerged, “new realities effectively came into being.”24 Hacking’s historical ontology

has provided a fertile ground for research. Scholars like Christopher Sellers, Michelle Murphy,

and Gabrielle Hecht have expanded Hacking’s emphasis—categories through which humans

might constitute themselves—to examine the conditions for the emergence and use of any

category, be it “occupational disease,” “sick building syndrome” or “nuclearity.”25

22 Ian Hacking, “The Looping Effects of Human Kinds,” in Causal Cognition: An Interdisciplinary Approach, ed. D. Sperber, D. Premack, and A. Premack (Oxford: Oxford University Press, 1995), 351-83. 23 Hacking’s earliest foray into historical ontology is, arguably, his work on the emergence of probabilistic understandings of certainty and uncertainty in the seventeenth century. Later publications have focused more directly on human categories that result in looping effects. See: Ian Hacking, The Emergence of Probability: A Philosophical Study of Early Ideas About Probability, Induction and Statistical Inference (New York: Cambridge University Press, 1975); Ian Hacking, Rewriting the Soul: Multiple Personality and the Sciences of Memory (Princeton, NJ: Princeton University Press, 1995); Ian Hacking, Mad Travelers: Reflections On the Reality of Transient Mental Illnesses (Charlottesville, VA: University Press of Virginia, 1998); Hacking, Historical Ontology, ch. 6. 24 Hacking, Historical Ontology, 101. 25 Christopher Sellers, Hazards of the Job: From Industrial Disease to Environmental Health Science (Chapel Hill, NC: University of North Carolina Press, 1997); Michelle Murphy, Sick Building Syndrome and the Problem of Uncertainty: Environmental Politics, Technoscience, and Women Workers (Durham, NC: Duke University Press,

11

Akin to Hacking, Lorraine Daston designates historical studies of concepts as “historical

epistemology.” For Daston, historical epistemology examines the fundamental categories that

made the production of scientific knowledge possible and how these have altered over time. It

is, according to Daston, “the history of the categories that structure our thought, pattern our

arguments and proofs, and certify our standards of explanation.”26 Daston identifies a particular

subset of historical epistemology as “applied metaphysics,” a label she applies to the approach

used by scholars in Biographies of Scientific Objects (2000).27 Contributors to that volume are concerned not with the “ethereal world of what is always and everywhere from a God's-eye- viewpoint” but with “the dynamic world of what emerges and disappears from the horizon of working scientists.”28 These authors focus on, “how whole domains of phenomena—dreams, atoms, monsters, culture, mortality, centers of gravity, value, cytoplasmic particles, the self, tuberculosis—come into being and pass away as objects of scientific inquiry.”29 Thus, applied

metaphysics is the study of how things become the objects of scientific inquiry, how they

become scientific objects. Scientific objects are historically constituted. They emerge at

particular periods of time as the result of the coalescence of various human interests,

experimental techniques, instruments, observational practices, and methods of analysis. Applied

metaphysics details, “how a heretofore unknown, ignored, or dispersed set of phenomena is

2006); Gabrielle Hecht, Being Nuclear: Africans and the Global Uranium Trade (Cambridge, MA: MIT Press, 2012). 26 Lorraine Daston, “Historical Epistemology,” in Questions of Evidence: Proof, Practice, and Persuasion across the Disciplines, eds. James Chandler, Arnold I. Davidson, and Harry Harootunian (Chicago: University of Chicago Press, 1994), 282. 27 Lorraine Daston, “Preface,” in Biographies of Scientific Objects, ed. Lorraine Daston (Chicago: University of Chicago Press, 2000), ix. 28 Lorraine Daston, “Introduction: The Coming into Being of Scientific Objects,” in Biographies of Scientific Objects, ed. Lorraine Daston (Chicago: University of Chicago Press, 2000), 1. 29 Ibid, 5.

12

transformed into a scientific object that can be observed and manipulated, that is capable of

theoretical ramifications and empirical surprises, and that coheres, at least for a time, as an

ontological entity.”30

There are strong methodological similarities between Daston’s applied metaphysics and

Hacking’s historical ontology.31 However, historical ontology provides a better label for this

dissertation’s approach. First, it has been explicitly taken up by more historians of science.32

Second, it more aptly captures the character of the Earth system as a new kind of thing in the

world. Historical ontologies focus on the historical alterations to, or the

emergence/disappearance of, certain categories as objects of knowledge. They examine how

ontologies grow, contract, or in some way change. The ESSC’s “Earth system” gave scientists a

common vocabulary to describe a systemic understanding of the planet that amorphously and

incrementally grew within Earth science communities from the 1960s into the early 1980s. With

the rise of the Earth system concept in the 1980s and its later entrenchment, there was a shift in

the Earth sciences and an expanded scientific ontology, an expanded understanding of what

could be studied scientifically, an expanded understanding of what counted as a scientific object

that included the whole Earth.

This dissertation examines the twentieth-century roots of the Earth system concept, its

development and promotion in the 1980s by a small group of scientists based in the US, and its

eventual adoption by practitioners across the Earth sciences. The physical thing that we live on

30 Ibid [emphasis added]. 31 Hacking distinguishes historical ontology from historical epistemology only to the extent that they have slightly different emphases. While Daston’s historical epistemology (or applied metaphysics) focuses on the coming into being of objects of scientific inquiry and the shifting meanings of the categories that make knowledge possible— with some of these objects previously existing as “quotidian objects”—Hacking states that historical ontology, “is concerned with objects or their effects which do not exist in any recognizable form until they are objects of scientific study.” See: Hacking, Historical Ontology, 11 [emphasis added]. 32 For an example of a sustained critique of historical epistemology and, by extension, applied metaphysics, see: Yves Gingras, “Naming Without Necessity: On the Genealogy and Uses of the Label ‘Historical Epistemology,’” Revue de Synthèse 131, no. 3 (2010): 439-54.

13

has (of course) always existed in a commonsense way for humans. But the planet took on a new

salience throughout the twentieth century for Earth scientists.33 It was no longer simply the

place where humans lived, but with the new tools of satellites and computer models it could

increasingly be the subject of more global scientific inquiry. The Earth system phrase represents

the culmination in this shift in salience, the transformation of a quotidian object (the planet) into

a scientific object (the Earth system). Hacking’s historical ontology, rather than Daston’s

applied metaphysics, better describes historical analyses—including the one offered here—that

focus on ontological shifts.34

The Earth system’s multi-level character differentiates it from the concepts examined in other historical ontologies. Its multifacetedness—where the Earth system can broadly mean one common thing to all Earth scientists though each discipline might differently construe its specific contours—allows for an unusual degree of local variety while simultaneously facilitating broad, interdisciplinary communication. The story of the Earth system concept is, as a result, a story of success and ubiquity. Earth scientists debated the details of specific Earth science research programs like Earth system science, but not the usefully vague Earth system concept itself which, largely without contest, filled the semantic void growing from the 1960s into the 1980s about how to concisely describe an interconnected planet. The Earth system had (and has) a broad meaning that was widely accepted amongst Earth science practitioners. To fully understand the Earth, they believed that its components should be studied as part of a singular, interconnected system through interdisciplinary collaboration. No Earth scientist disputed this

“low resolution” meaning of the Earth system. This places the concept at odds with the frequently contested categories presented by Hacking, Sellers, Murphy, Hecht, and Daston, as

33 Daston, “Introduction,” 6-9. 34 Hacking describes ontology as a “branch” of metaphysics. See: Hacking, Historical Ontology, 1.

14

well as with other history of science scholarship examining theoretical entities like the electron

or microbes.35

For the Earth system, contestation arises from the ways it is differently defined by specialists in the various Earth science disciplines, though such differences are not necessarily sources of friction. Different Earth scientists will draw different spatial and temporal boundaries based on research interests. Atmospheric physicists might define the Earth system to focus on

shorter term processes in the atmosphere and oceans, like wind patterns or precipitation.

Geologists, however, recognize that atmospheric physicists exclude their longer term processes

for practical purposes only. Atmospheric physicists ignore long-term geological processes not because they perceive those processes to be unscientific, but simply because the world’s vast complexity requires, for all practical purposes, more focused study. These specialist definitional differences themselves do not generate friction or controversy. Trouble arises when real resources are at stake, when scientists are at risk of having their research areas, and thereby their research funding, marginalized. But this trouble arises from competition for scarce resources, not with the Earth system concept itself (chapter three).

Sources

This historical ontology of the Earth system relies predominantly on archival material, supplemented with select secondary material and published oral histories. The most significant and unique archival resource comes from ESSC chairman Francis Bretherton’s unprocessed personal records housed at the National Center for Atmospheric Research (NCAR). Up to now, the ESSC has received little attention in the extant historiography that has heretofore focused

35 Theodore Arabatzis, Representing Electrons: A Biographical Approach to Theoretical Entities (Chicago: University of Chicago Press, 2006); Bruno Latour, The Pasteurization of France, trans. Alan Sheridan and John Law (Cambridge, MA: Harvard University Press, 1988).

15 predominantly on NASA’s longue durée Earth science research activities or NASA’s development of the Earth Observing System (EOS) satellites.36 This research has not devoted more than a few paragraphs or pages to the specific work done by the ESSC. This lacuna may in part be traceable to the available sources on the ESSC. Until recently, these consisted primarily of a single small file housed at NASA’s Historical Reference Collection (HRC) and the ESSC’s two published reports (the Overview and the Closer View).37 However, these sources provide only a narrow and fragmentary picture of the ESSC’s work.

Bretherton’s files offer an unprecedented view of the activities of ESSC members and external collaborators. They facilitate the construction of a more detailed and nuanced portrait of the group than has been previously possible. These documents—which include formal correspondence, rough notes, sketches, diagrams, photographs, reports, and early electronic mail communications between committee members—demonstrate the great efforts undertaken by

ESSC members to spread their ideas and garner support. Analyzing these documents helps explain why the ESSC’s Earth system concept has become so dominant in the Earth sciences today. The electronic mail correspondence in particular provides unprecedented access to the

36 Erik M. Conway, Atmospheric Science at NASA: A History (Baltimore: Johns Hopkins University Press, 2008); Erik M. Conway, “Bringing NASA Back to Earth: A Search for Relevance during the Cold War,” in Science and Technology in the Global Cold War, eds. Naomi Oreskes and John Krige (Cambridge, MA: MIT Press, 2014), 251– 72; Edward S. Goldstein, “NASA’s Earth Science Program: The Bureaucratic Struggles of the Space Agency’s Mission to Planet Earth” (PhD dissertation, George Washington University, 2007); W. Henry Lambright, “Administrative Entrepreneurship and Space Technology: The Ups and Downs of ‘Mission to Planet Earth,’” Public Administration Review 54, no. 2 (Mar./Apr. 1994): 97–104; W. Henry Lambright, “The Political Construction of Space Satellite Technology,” Science, Technology, and Human Values 19, no. 1 (Winter 1994): 47–69; Richard B. Leshner, “The Evolution of the NASA Earth Observing System: A Case Study in Policy and Project Formulation” (PhD dissertation, George Washington University, 2007); John H. McElroy and Ray A. Williamson, “The Evolution of Earth Science Research from Space: NASA’s Earth Observing System,” in Exploring the Unknown: Selected Documents in the History of the US Civil Space Program: Volume VI: Space and Earth Science, eds. John M. Logsdon, et al. (Washington, DC: NASA, 2004), 441-73; Richard B. Leshner and Thor Hogan, The View From Space: NASA's Evolving Struggle to Understand Our Home Planet (Lawrence, KS: University Press of Kansas, 2019). 37 Box 18042, NASA Historical Reference Collection (HRC), Washington, DC; Earth System Sciences Committee (ESSC), Earth System Science: A Program For Global Change: Overview (Washington, DC: NASA, 1986); Earth System Sciences Committee (ESSC), Earth System Science: A Program For Global Change: A Closer View (Washington, DC: NASA, 1988).

16

seemingly uncensored opinions, frustrations, concerns, and excitements of the actors as they

existed at that time. This important resource is supplemented by archival material from NASA’s

HRC and the American Geophysical Union (AGU) Records housed at the American Institute of

Physics’ (AIP) Niels Bohr Library and Archives.

MAKING THE GLOBAL ENVIRONMENT

With its Earth system science research program, the ESSC sought to provide a framework for

more global and interdisciplinary studies of the planet. While Earth system science failed to gain

widespread support, the scope of the ESSC’s work was unequivocally global, and its Earth

system phrase spread quickly and profusely as the common way for Earth scientists to refer to

the entire interconnected planet. This dissertation, therefore, is situated at the intersection of a

number of different threads in the history of the Earth and environmental sciences broadly

captured by what historian of science Lino Camprubí calls a, “new historiography of the global

environment.”38 Of course, as Paul Edwards suggests, “The idea of ‘the world’ is probably as old as language, with as many meanings and connotations as there are cultures.”39 However, by the late nineteenth century, increasing numbers of Earth scientists focused their attention on larger scale studies and planetary theorizations, aided in the mid-twentieth century by new technologies and techniques that provided new sources of empirical data and analysis capabilities. Camprubí argues that, though there is a long and rich history of thinking globally about Earth, only very recently has the “global environment” become a focus for scholarship.40

38 Lino Camprubí, “The Invention of the Global Environment,” Historical Studies in the Natural Sciences 46, no. 2 (2016): 249. 39 Paul N. Edwards, “The World in a Machine: Origins and Impacts of Early Computerized Global Systems Models.” in Systems, Experts, and : The Systems Approach in Management and Engineering, World War II and After, eds. Agatha C. Hughes, Thomas P. Hughes (Cambridge, MA: MIT Press, 2000), 221. 40 Camprubí, “The Invention of the Global Environment,” 243-4.

17

Camprubí and Philipp Lehmann describe the “global environment” as an idea that begins to suffuse the many sprawling Earth and disciplines only in the latter part of the nineteenth century. It emerges as a “commitment” to the development of global representations of the planet, be they global models and datasets, photographs, metaphors, or concepts.41 Rather than treat the global environment as a self-evident or necessary category, scholars working in this area interrogate and illuminate the situated, local contexts within which the global environment (or more properly global environments) has been constructed.42

Scholarship on the global environment is still nascent, with a shifting and expanding topography

on a number of different research fronts. This dissertation contributes to the growing body of

work at the intersection of three key areas of research in the history of the global environment:

issues of scale, views of the Earth from above, and Cold War Earth sciences.

Given that “global” is in the name, it is not surprising that issues of scale emerge

frequently in the global environment historiography. Camprubí and Lehmann describe the issue

as a persistent tension in the construction of the global environment with respect to experience.

How can humans, necessarily linked to a local context, actually develop an empirical conception

of the global environment?43 This is arguably a reframing of the Humean problem of induction:

the problem of how particular observed cases project to more general unobserved cases.44

Rather than remain stuck in a potentially philosophically intractable problem, scholars use

41 Lino Camprubí and Philipp Lehmann, Program for “Experiencing the Global Environment,” Max Planck Institute for the History of Science, 4-6 February 2016, accessed 15 Dec 2019, https://www.mpiwg- berlin.mpg.de/sites/default/files/migrated/program_experiencing_global_environment_1.pdf. 42 Lino Camprubí and Philipp Lehmann, “The Scales of Experience: Introduction to the Special Issue Experiencing the global environment,” Studies in History and Philosophy of Science Part A 70 (2018): 2-3. 43 Camprubí and Lehmann, “The Scales of Experience,” 1; Donna J. Haraway, “Situated Knowledges: The Science Question in Feminism and the Privilege of Partial Perspective,” in Feminism and Science, eds. Evelyn Fox Keller and Helen E. Longino (New York: Oxford University Press, 1996), 249-63. 44 David Hume, An Enquiry Concerning Human Understanding, ed. Tom L. Beauchamp (Toronto: Oxford University Press, 1999), 108-18.

18

historical case studies to illuminate how actors built up conceptions of the global environment

based on localized experiences.45 They do not ask whether or not this transformation is

philosophically legitimate. Instead, they focus on how historical actors have made the leap via a

variety of mediations and techniques. Contributors to the special issue on “Experiencing the

Global Environment” in Studies in History and Philosophy of Science (2018) examine a variety

of methods, including: measurements of human bodily exposures to map environmental

pollutants, non-human animal perception of seismic instabilities in tectonic plates, standardized

local observational techniques to develop global views of weather and climate, visual diagrams

to depict global ocean currents, rhetorical strategies to promote global geomorphology studies,

and the use of transportation and communication networks to map distributions of birds and

mammals.46 Earlier work in the history of , oceanography, and the environmental sciences presaged these issues of scale.47 Taking a slightly different tack, Deborah Coen rejects

the idea that there are fixed, final, a priori scales that differentiate local and global contexts.

What makes something a “human” scale varies depending on factors like lifespan, degrees of

mobility, access to communication and transportation networks, and ancestral memory (or lack

thereof). Scales are not fixed in advance, but are historical outcomes and should, therefore, be interrogated and historicized.48

45 From the perspective of deductive logical reasoning, the problem of induction is philosophically intractable. However, scholars have attempted to reframe the problem in order to curtail its significance. See: Nelson Goodman, Fact, Fiction, and Forecast (New York: The Bobbs-Merrill Company, 1965). 46 “Special Issue: Experiencing the global environment,” Studies in History and Philosophy of Science Part A 70 (2018): 1-86. 47 James Rodger Fleming, Vladimir Janković, and Deborah R. Coen, eds., Intimate Universality: Local and Global Themes in the History of Weather and Climate (Sagamore Beach, MA: Science History Publications/USA, 2006); Tiffany C. Vance and Ronald E. Doel, “Graphical Methods and Cold War Scientific Practice: The Stommel Diagram’s Intriguing Journey from the Physical to the Biological Environmental Sciences,” Historical Studies in the Natural Sciences 40, no. 1 (2010): 1-47. 48 Deborah R. Coen, “Big is a Thing of the Past: Climate Change and Methodology in the History of Ideas,” Journal of the History of Ideas 77, no. 2 (Apr. 2016): 305-21. Coen takes up her own call in Climate in Motion (2018), where she examines the history of dynamic climatology in the vast and heterogeneous Hapsburg Empire in the

19

Issues of scale play an important role in this dissertation. Earth scientists here were not

doing science, so this dissertation does not trace how scientists translate local practices and

knowledge creation into global understandings in the same way as other scholars have in the

historiography. But ESSC members were attempting to construct a new global interdisciplinary

Earth system science research program that would reorganize the formerly disparate and

compartmentalized Earth sciences into something more cohesive. They failed in that objective,

but their broad framing of the interconnected character of Earth processes and their specific

name for these interconnections—the Earth system—attained widespread, indeed global, usage.

This dissertation also contributes to the politics of scale by examining the implications of

delineating temporal scales. As Coen argues, there was no a priori choice for timescale

prioritization. ESSC members fully recognized that scale was a construct that could differ

according to human interests and concerns, so much so that some members even wrote and

shared jokes about this issue.49 Certain ESSC members justified the prioritization of timescales

decades to centuries over longer ones for practical reasons (not every timescale could be

practically incorporated into an Earth system model of the entire planet) and the environmental

urgency of addressing processes occurring on timescales of a human lifespan. However, this

decision created considerable backlash from researchers focused on longer term processes that

were effectively marginalized by the decision. This contributed to the lukewarm reception Earth

system science received at the time.

nineteenth century whereby climatology was rendered “global science.” Deborah R. Coen, Climate in Motion: Science, Empire, and the Problem of Scale (Chicago: University of Chicago Press, 2018). 49 This recognition forms the humourous backdrop for the Dynos joke described in chapter three.

20

The second area of global environmental historiography this dissertation intersects with is

a different issue related to scale, what I call views of the Earth from above. Scholars working in

this area examine the ways that perspectives of the Earth beyond its surface—be they from

airplanes, satellites, other spacecraft, or simply metaphorical views—provide new ways of

conceptualizing the planet. The emphasis on plurality is crucial. Scholars like Denis Cosgrove,

Sheila Jasanoff, and Robert Pool stress—using the Earthrise and Blue Marble images of the

Earth from space taken by Apollo astronauts—that there is no single, given, universal,

unambiguous perspective necessitated by these overviews. Different groups with different

concerns and agendas may differently interpret these images or metaphors. Further, it takes

work to establish common ways of “seeing.”50 As Jeanne Haffner argues in The View From

Above (2013), technologies like airplanes often played crucial roles in reconfiguring environmental and social imaginaries by facilitating a “detached” view of the planet from above that allowed the detection of patterns and relationships not observable from the surface.51 Sabine

Höhler examines how these views from above, made possible by the space technologies of the

Cold War, contributed to the rise of the “Spaceship Earth” metaphor in the 1960s. The metaphor

identified the planet as a closed system with finite boundaries threatened by human actions.

Spaceship Earth mixed environmental concerns with hopes for techno-optimistic solutions.52

Perrin Selcer adds an international institutional element to the Spaceship Earth metaphor by

investigating how United Nations specialized agencies attempted to establish a post-World War

50 Denis Cosgrove, “Contested Global Visions: One-World, Whole-Earth, and the Apollo Space Photographs,” Annals of the Association of American Geographers 84, no. 2 (1994): 270-94; Denis Cosgrove, Apollo’s Eye: A Cartographic Genealogy of the Earth in the Western Imagination (Baltimore, MD: Johns Hopkins Press, 2001); Sheila Jasanoff, “Image and Consciousness: The Formation of Global Environmental Consciousness,” in Changing the Atmosphere: Expert Knowledge and Environmental Governance, eds. Clark A. Miller and Paul N. Edwards (Cambridge, MA: MIT Press, 2001), 309-38; Robert Poole, Earthrise: How Man First Saw the Earth (New Haven, CT: Yale University Press, 2008). 51 Jeanne Haffner, The View From Above: The Science of Social Space (Cambridge, MA: MIT Press, 2013). 52 Höhler, Spaceship Earth.

21

II international liberal order by transforming local and regional concerns and capabilities into a

legible global environment, into an interconnected “Spaceship Earth.”53

This dissertation highlights the ways that views from above facilitated an understanding

of an interconnected, finite planet. However, this dissertation studies neither photographs of the

Earth from space nor the Spaceship Earth metaphor. The story here recounts the global

conception of the planet that emerged amongst practicing Earth and environmental scientists. A

group of scientists attempted to construct a global research program heavily reliant on satellites

to study the Earth as a system. While this program failed to gain traction, the group’s systemic

conception of the planet—the Earth system—did. This dissertation examines a particular “view”

from above. It is a view not from a satellite but a model of the earth from “above.” The model

was visually expressed using a “wiring diagram” that structurally depicted Earth as a number of

interconnecting subsystems comprising the air, water, land, ice, and biota (Figure 3.1). The

Earth system wiring diagram did not, as the ESSC hoped, provide the starting point for an

interdisciplinary Earth system science research program. Nevertheless, this particular view from

above did present a broad and immediately understandable conceptual framework to understand

how the various Earth science disciplines connected together in their complementary studies of

an interconnected planet.

The new technologies and new funding made available by the Cold War not only affected images and metaphors of the planet. They also provided new research opportunities in the Earth sciences. This dissertation lastly intersects with literature focused on a third key area in the historiography of the global environment, Cold War Earth science. Camprubí identifies the Cold

War as an important turning point in the nascent but growing historiography of the global

53 Perrin Selcer, The Postwar Origins of the Global Environment: How the United Nations Built Spaceship Earth (New York: Press, 2018).

22

environment, particularly for Earth and environmental science scholarship.54 Many histories of

Cold War Earth and environmental science focus on the effects of greatly increased military

funding for these disciplines in the US. With new or increasingly important and global

technologies like submarines, aircraft, intercontinental ballistic missiles, and satellites came the

parallel military need to increase knowledge of the planet’s total environment where these

technologies operated, be it on land, in the oceans, in the air, or in near-space. Scholars like

Ronald Doel, Kristine Harper, Jacob Darwin Hamblin, David Munns, the contributors to Naomi

Oreskes and John Krige’s Science and Technology in the Global Cold War (2014), and

contributors to J.R. McNeil and Corinna Unger’s Environmental Histories of the Cold War

examine scientists’ activities during the period to understand the concrete effects of military

patronage and Cold War geopolitics on Earth and environmental science enterprises. The effects

often manifested as emphases on the physical over biological sciences and attempts to control or

weaponize the environment for military advantage.55

Work by Joel Hagen, Donald Worster, Hamblin, and David DeVorkin emphasize the

mutually beneficial alignment of military and scientific research interests during this period.56

Military interests drove much early post-WWII Earth science research. But as Paul Edwards,

Joseph Masco, and contributors to Simone Turchetti and Peder Roberts’ The Surveillance

54 Camprubí, “The Invention of the Global Environment,” 246. 55 Ronald E. Doel, “Constituting the Postwar Earth Sciences: The Military’s Influence on the Environmental Sciences in the USA after 1945,” Social Studies of Science 35, no. 5 (Oct. 2003): 635-66; Kristine C. Harper, Make It Rain: State Control of the Atmosphere in Twentieth-Century America (Chicago: University of Chicago Press, 2017); Jacob Darwin Hamblin, Arming Mother Nature: The Birth of Catastrophic (Toronto: Oxford University Press, 2013); David P. Munns, Engineering the Environment: Phytotrons and the Quest for Climate Control in the Cold War (Pittsburgh: Pittsburgh University Press, 2017); Naomi Oreskes and John Krige, eds., Science and Technology in the Global Cold War (Cambridge, MA: MIT Press, 2014); J.R. McNeil and Corinna R. Unger, eds., Environmental Histories of the Cold War (New York: Cambridge University Press, 2010). 56 Joel B. Hagen, An Entangled Bank: The Origins of Ecosystem (New Brunswick, NJ: Rutgers University Press, 1992); Donald Worster, Nature’s Economy: A History of Ecological Ideas, second edition (New York: Cambridge University Press, 1994); Jacob Darwin Hamblin, Oceanographers and the Cold War: Disciples of Marine Science (Seattle: University of Washington Press, 2005); David H. DeVorkin, Science with a Vengeance: How the Military Created the US Space Sciences after World War II (New York: Springer-Verlag, 1992).

23

Imperative (2014) argue, military weapons and surveillance technologies—notably satellites and

nuclear weapons research, which allowed scientists to study the large-scale movement of radioactive elements in the atmosphere, oceans, and biota—also provided new ways to research and understand the Earth as an interconnected, and often fragile, global environment.57 Hamblin

shows that US officials during the Cold War sought to both weaponize this environmental

knowledge of a fragile and interconnected planet through techniques like radiological

contamination, biological weapons, and weather control, and to protect its potentially vulnerable

population against this “environmental warfare,” providing a foundation for global

environmental concerns.58

The relationship between the military and Earth science research during the Cold War has

been amply demonstrated by scholars, but this is only one component of a broader Cold War

history. The singular phrase, the “Cold War,” tends to obscure a motley and multifaceted period

that saw fluctuations in superpower tensions and military funding for basic research. There is no

singular “Cold War” that provides the context for scientific and technological activities. In terms

of military patronage for the Earth sciences, this flow lessened significantly in the latter decades

of the Cold War. Audra Wolfe, in Competing With the Soviets (2013), identifies the Pentagon’s

mid-1960s “Project Hindsight” as a major turning point. The report concluded that military

value for “undirected science” could not be easily demonstrated. This helped shatter post-WWII

technoscientific optimism and led to the passage of the Military Authorization Act and Mansfield

57 Paul N. Edwards, A Vast Machine: Computer Models, Climate Data, and the Politics of Global Warming (Cambridge, MA: MIT Press, 2010), ch. 8; Joseph Masco, “Terraforming Planet Earth: The Age of Fallout,” in The Politics of Globality Since 1945: Assembling the Planet, eds. Rens van Munster and Casper Sylvest (New York: Routledge, 2016), 44-70; Joseph Masco, “The Age of Fallout,” History of the Present 5, no. 2 (Fall 2015): 137-68; Joseph Masco, “Bad Weather: On Planetary Crisis,” Social Studies of Science 40, no. 1 (Feb. 2010): 7-40; Simone Turchetti and Peder Roberts, eds., The Surveillance Imperative: Geosciences during the Cold War and Beyond (New York: Palgrave MacMillan, 2014). 58 Hamblin, Arming Mother Nature.

24

Amendment in 1970, which dramatically restricted military funding for civilian projects.

Subsequent scientific research supported with government dollars increasingly required a

demonstration of practical value.59

The period after 1970, therefore, provides a much different context for US scientific

research in general and Earth science research in particular. Given its close chronological

proximity, historiography on this later Cold War period is still sparse but it is growing. Rachel

Rothschild’s Poisonous Skies (2019) provides a recent example. Rothschild studies Cold War

acid rain research in the 1970s and 1980s—where military sponsorship was peripheral at most—

to show how environments were reconceived from discrete and isolated to regionally and

globally interconnected.60 Kai Hünemörder argues that transnational environmental pollution

concerns in the late 1960s and 1970s were actively deployed in the service of Cold War détente

foreign policies. Hünemörder shows how East and West German officials attempted to use the

rhetoric of “environmental protection” and “environmental crisis” to foster political cooperation

across the Iron Curtain.61 Particularly relevant is the work by Erik Conway, Neil Maher, and W.

Henry Lambright that focuses on the increasing importance of Earth science research at NASA

after 1970. This research exemplified later Cold War global science that lacked strong military

59 Audra J. Wolfe, Competing With the Soviets: Science, Technology, and the State in Cold War America (Baltimore, MD: The Johns Hopkins University Press, 2013), 120-2. As a percentage of the total US federal budget, federal R&D funding peaked in 1965 at 11.7%. In 2018, R&D expenditures were only 2.8% of the total budget. See: American Association for the Advancement of Science, “Historical Trends in Federal R&D - R&D as Percentage of the Total Federal Budget, 1962-2018,” accessed 19 Dec 2019, https://www.aaas.org/programs/r-d- budget-and-policy/historical-trends-federal-rd. 60 Rachel Emma Rothschild, Poisonous Skies: Acid Rain and the of Pollution (Chicago: University of Chicago Press, 2019). 61 Kai Hünemörder, “Environmental Crisis and Soft Politics: Détente and the Global Environment, 1968–1975,” in Environmental Histories of the Cold War, eds. J.R. McNeil and Corinna R. Unger (New York: Cambridge University Press, 2010), 257-76.

25

support, though it still relied on technologies like satellites and computers with some military

provenance.62

Paul Edwards offers one of the most comprehensive studies of how Earth scientists

construct global knowledge. Edwards’ A Vast Machine (2010) targets the activities of climate

scientists and meteorologists in their development of global observational networks to “make

global data” and their subsequent computer modeling efforts to transform those global datasets

into global knowledge, what Edwards calls “making data global.”63 Edwards’ analysis spans not

only the entire Cold War period but also the early days of meteorological data collection in the

nineteenth century and contemporary climate modeling. He traces primarily US efforts to

illuminate the knowledge infrastructure that underpins current understandings of global climate

change. It is undeniable that there was military involvement in the major technologies—Earth

observing satellites and digital computers—that facilitated global datasets and analyses. Indeed,

Edwards meticulously details one aspect of this relationship in The Closed World (1996), where

he traces the history of US military support for and use of computers post-WWII and how computers fueled a “closed” understanding of the planet which, in turn, supported military battlefield imaginaries of total control.64 In “The World in a Machine” (2000), Edwards further

elaborates on these roots with respect to computer modeling of global systems. Edwards

attributes these models, along with satellites (which also have direct military origins), as

providing the necessary technological prerequisites for viewing the planet as a system.65

62 Conway, Atmospheric Science at NASA; Conway, “Bringing NASA Back to Earth,”; Neil M. Maher, Apollo in the Age of Aquarius (Cambridge, MA: Harvard University Press, 2017); W. Henry Lambright, NASA and the Environment: The Case of Ozone Depletion (Washington, DC: NASA, 2005). 63 Edwards, A Vast Machine, xv, ch. 10. 64 Paul N. Edwards, The Closed World: Computers and the Politics of Discourse in Cold War America (Cambridge, MA: MIT Press, 1996). 65 Paul N. Edwards, “The World in a Machine: Origins and Impacts of Early Computerized Global Systems Models,” in Systems, Experts, and Computers: The Systems Approach in Management and Engineering, World War

26

However, Edwards never suggests that global climate knowledge or global modeling more broadly are simply military concerns cloaked in civilian clothing, nor does he suggest that military origins somehow taint subsequent scientific efforts. Much research in these fields occurred post-1970, during the period where military funding for Earth science research became significantly less pronounced.

This dissertation builds on Edwards’ argument that Cold War technologies like satellites and computer modeling provided the conditions for the possibility of viewing the Earth as a system. These technologies, no doubt, emerged out of Cold War competition and surveillance imperatives, as well as the desire for real-time command and control. They arguably provided necessary technical capabilities, but by themselves these were not sufficient to transform the

Earth into a scientific object identified by the ubiquitous phrase “Earth system.” This further step required the coordinating work of a select group of Earth and environmental scientists working within the institutional mechanisms of NASA. While it is historically situated in the

Cold War, this dissertation breaks with much of the Cold War Earth science literature to the extent that military interest and funding is largely absent, apart from the technological legacies of satellites and computers. Government support for Earth science was still important, but it was much more restricted after 1970 when military involvement in scientific research became comparatively reduced. NASA itself faced significant funding declines by the end of the 1960s and in the early 1970s.66 Hence, in the early 1980s, NASA needed to form a quasi-external

II and After, eds. Agatha C. Hughes and Thomas P. Hughes (Cambridge, MA: MIT Press, 2000): 221-53. See also: Turchetti and Peder Roberts, eds., The Surveillance Imperative. 66 As Conway and Maher argue, post-1972 (the end of the Apollo Program) NASA had to develop new justifications for any government expenditures, resulting in the agency’s turn “back to Earth.” See: Conway, “Bringing NASA Back to Earth”; Maher, Apollo.

27

committee to develop a large-scale research program in which the agency would (hopefully) play

a central role. This program was intended to garner widespread support from both scientific and

political communities. After 1970, government funding for civilian scientific research could not

be assumed.

There are multiple ways humans have historically constituted a global environment, as a

product of cultural, social, political, military, aesthetic, technological, or scientific interests and

concerns.67 This dissertation focuses on the sociological, scientific, technological, and

institutional factors emerging in the 1980s in the Earth and environmental sciences that

facilitated a particular conception of the global environment. Many scholars have examined the

practices whereby different Earth science communities constructed knowledge on the planet,

whether more local or more global in scale. However, no book-length scholarship has examined

how a new common conception of the global environment emerged in the 1980s and took hold in

the various Earth science disciplines. This dissertation fills this space by examining the specific

roots of this particular conception of the global environment—the Earth system—that provided a

singular way of referring to the Earth as an interconnected planet amenable to scientific inquiry.

There are numerous, more colloquial ways that Earth scientists might refer to Earth, but when

they refer to the planet specifically as a Dastonian scientific object, they use the phrase “Earth system” frequently and unreflectively. This dissertation studies the early conditions and activities that made this possible.

67 As a result, Camprubí and Lehmann suggest that this area should more properly be called a historiography of “global environments.” See: Camprubí and Lehmann, “The Scales of Experience,” 2.

28

CHAPTER OUTLINES

The specific phrase—the Earth system—was a product of the ESSC. However, this phrase and what it referred to were not developed and promoted ex nihilo. That it could spread so readily

amongst Earth scientists in the 1980s suggests that the phrase was apt to capture sentiments

already present in these communities. Using a mixture of published primary and secondary

material, chapter one examines these early conditions. It adapts Leo Marx’s concept of a

semantic void to analyze the conditions in the Earth and environmental sciences from the 1960s

into the 1980s that made them particularly receptive to the phrase “Earth system.” New

technologies and techniques like Earth observing satellites, digital computers, and systems

thinking provided new capacities post-World War II for Earth scientists to collect and analyze

global datasets in comprehensive ways that were unprecedented. Developing global theories was

not a new twentieth century phenomenon—as the chapter’s examples from Alexander von

Humboldt and Vladimir Vernadsky indicate—but these earlier theories were largely speculative

and lacked substantial empirical evidence. Beginning in the 1960s, Earth scientists turned to

newly available satellites and computer modeling to construct global knowledge. These tools

allowed for the empirical study of the entire Earth for the first time. They also supported a view

of the planet as an interconnected whole, as a system with interlinking subsystems that required

interdisciplinary collaboration to produce robust planetary knowledge. A new view of an

interconnected planet emerged, but there was no singular, common way to refer to this new

understanding. Thus, these new capacities arguably created a semantic void in which there was

no current and broadly accepted single phrase that adequately or finally described the Earth as a system. This is the void that would later be filled by the ESSC’s “Earth system.” Chapter one ends with three illustrative examples from the 1970s and early 1980s: an institution (NCAR), an international research program (the Global Atmospheric Research Program), and a scientific

29

hypothesis (James Lovelock’s Gaia hypothesis). These examples point to and emerged from the global orientation facilitated by satellites and computer modeling as well as the growing interest in interdisciplinary studies of the Earth.

Chapter one depicts the conditions that created the semantic void in the Earth sciences from the 1960s into the 1980s. However, not any phrase could fill this void, as the chapter’s

Gaia example shows. In addition to its useful vagueness, what made the Earth system appealing was the ESSC’s extensive preparatory and promotional work, as part of their proposed Earth system science research program. These efforts gave the Earth system concept wide exposure.

But the ESSC was not the first group to attempt to transform the whole Earth into an object of scientific inquiry. NASA made an earlier attempt in 1982 with an initiative called Global

Habitability. Using published primary and secondary sources, along with archival documents from NASA’s HRC, chapter two describes this earlier attempt, one of the first major attempts by an institution to formulate a large-scale research program. It was based on the nascent and growing technological capacities afforded by satellites and computer modeling, and would treat the entire Earth as a system, as a scientific object. If this research program had been widely adopted, it is possible that Earth scientists might today refer to an interconnected planet as the

“global habitat” rather than “Earth system.” Instead, Global Habitability not only failed to gain

extensive support, but it received a hostile reception when introduced at the United Nations

Conference on the Exploration and Peaceful Uses of Outer Space in 1982 (UNISPACE ‘82).

This chapter examines some of the reasons for this negative reaction, largely centering on the

geopolitically sensitive issue of satellite data collection for a proposed initiative that incorporated

no clear data policy. It investigates the unique historical moment in which Global Habitability

was both conceived and rejected to illuminate the challenges involved in transforming the entire

Earth into a scientific object.

30

Though Global Habitability failed to garner widespread support, NASA officials continued their push to develop a global research program reliant on Earth observing satellites and computer modeling, thus continuing the move to transform the entire Earth into a scientific object in one large-scale research program. To do so, NASA’s Advisory Council formed the

ESSC, comprised of representatives from various Earth and environmental science disciplines, to formulate a comprehensive program to study the entire Earth as a system with interconnected subsystems in the air, lands, waters, and biota. Learning a number of lessons from Global

Habitability, as well as fortuitously benefitting from a more congenial political climate for satellite data collection, chapter three uses never-before-analyzed ESSC archival material to outline the committee’s painstaking work from 1983 to 1988 as it carefully crafted a detailed research program they called Earth system science. Today, Earth system science is a vague phrase that broadly denotes scientists conducting interdisciplinary studies of how the Earth’s components function together as a “whole system.”68 But in the 1980s, Earth system science referred to the specific research program formulated by the ESSC that prioritized the study of processes occurring on timescales of decades to centuries and data collection from Earth observing satellites. As part of its program, the ESSC described an interconnected planet that required interdisciplinary study, calling this view of the planet the “Earth system.” The Earth system was depicted in a single wiring diagram that showed, at a glance, how the planets’ components connected with each other and how the different Earth science disciplines could collaborate to produce global knowledge. The ESSC’s program was not wholly uncontroversial, however. Divides emerged as a result of the vastly different timescales used in different disciplines. For some, the ESSC controversially chose to focus on shorter timescales of decades

68 See: Tim Lenton, Earth System Science: A Very Short Introduction (New York: Oxford University Press, 2016).

31 to centuries, driven by the hope to eventually model the entire Earth system in a single computer model. This would not be practical if all timescales—from the very long to the infinitesimally short—were included. These divides help explain the lukewarm reception for the Earth system science research program from many outside the ESSC, even while the Earth system concept gained quick and widespread adoption.

One important lesson the ESSC learned from the Global Habitability stumble was the importance of building consensus for such a large-scale research program. Members did this by incorporating as many external scientists as possible into the program’s development process and by devoting extensive resources to the promotion of the Earth system science research program.

Chapter four uses the ESSC’s unprocessed archival material to show these far-reaching consensus-building activities. The ESSC envisioned Earth system science to be a comprehensive, Apollo-esque program requiring enormous resources for the development of sophisticated Earth observing satellites and computer models and information storage systems, and the coordination of expertise from science and engineering communities. It would be an expensive and time-consuming endeavour that would not be possible without widespread scientific, engineering, popular, and political support. The ESSC, therefore, devoted significant efforts to building this broad backing. The Global Habitability failure was always close to mind as a warning about the pitfalls of ignoring the need to build wide-ranging support for a major research program. The group disseminated and promoted their ideas through a wide variety of mechanisms and strategies that included collaborative report preparations and review processes, an emphasis on a clear “branding” for all ESSC material, and a variety of strategies to spread awareness of Earth system science via highly polished publications, popular news articles, a press conference, and even Earth system science products like posters and t-shirts. All these promotional activities help account for why the specific phrase “Earth system” won out as the

32

way to describe the planet as a scientific object, while other phrases—phrases like “global

system” or “climatic system” or “the coupled land-ocean-atmosphere system”—fell into disuse.

The ESSC may not have intended to entrench the “Earth system,” but one of the most lasting effects of the group’s work was the way that the phrase rapidly spread beyond its confines to rapidly become the predominant way that Earth scientists referred to an interconnected planet requiring interdisciplinary collaboration. Chapter five uses archival material from a prominent

US Earth science organization—the American Geophysical Union (AGU)—to illustrate how the

Earth system concept was readily adopted by many external Earth scientists, even those not favourably disposed towards the ESSC’s specific Earth system science research program. AGU members were drawn to the comprehensive nature of the Earth system phrase not because they wanted to build a single model of the entire system but because they wanted to comprehensively study the entire Earth and its environs, from its inner core to the near space environment and even interactions with the . The Earth system phrase proved useful to various Earth scientists not because it meant something narrow and specific but rather because it was vague. It served as what Susan Leigh Star calls a boundary object, something that facilitates broad communication and collaboration despite lack of consensus at the local level. A boundary object provides narrow meaning at the specialist level while at the same time maintaining enough interpretive flexibility to transcend specialist uses to provide common meaning amongst different groups.

Different Earth science disciplines defined the “Earth system” differently (by either incorporating or excluding processes occurring on different timescales). Nevertheless, Earth scientists broadly agreed that the Earth should be studied as a system. Issues arose only over where to draw the boundaries of the system and how it might best be studied (for instance, in-situ

versus satellite instruments). Scientists disagreed about certain details, but they agreed that the

Earth was comprised of interconnected components that required interdisciplinary research. The

33

AGU case study provides an example of the efficaciousness of at least part of the ESSC’s development and promotional efforts. It shows how knowledge and concepts flow between different practitioners and institutions and scientific communities. AGU activities in the late

1980s demonstrate the organization’s receptiveness to the Earth system as a unifying phrase to describe the planet. The case study gestures towards the concept’s later entrenchment and near ubiquity in the Earth sciences.

The dissertation ends by examining a potential implication of the Earth system’s current entrenchment and ubiquity as it relates to the present climate emergency. Since it has become so naturalized, and since it can—though need not—be interpreted in a reductionistic way that supports notions of total knowledge and control of the planet, the Earth system could be illegitimately used to justify or support major geoengineering projects.

Chapter 1 The Emerging Threads: US Earth Sciences from the 1960s to the 1980s

INTRODUCTION

Since about 1986, Earth scientists have used the phrase “Earth system” to refer to the planet as

an interconnected system comprised of the atmosphere, oceans, land, ice, and biosphere. This

usage coincides precisely with the work of the Earth System Sciences Committee (ESSC).

Through the ESSC’s development and promotional work (chapters three and four), the Earth

system concept gained widespread traction and became the predominant way that Earth and

environmental scientists refer to the Earth as a scientific object. While the ESSC provided the

specific phrase, the ideas that the phrase gestured at grew out of certain conditions and

capabilities in Earth science communities that emerged in the 1960s and increased in intensity

throughout the 1970s and early 1980s. These conditions created, borrowing from the historian of

technology Leo Marx, a “semantic void.” This “void” was a lack of a generally accepted term to

refer to an Earth that scientists were increasingly able to study in more global, interdisciplinary,

and interconnected ways. With new technologies—notably satellites, computer modeling, and

systems thinking—scientists believed that they possessed the tools to construct global knowledge

of the entire planet. The Earth could now become, in Lorraine Daston’s words, a scientific

object.1

This chapter broadly examines the conditions in the Earth sciences that contributed to the

semantic void that would later be filled by the “Earth system” concept. The chapter begins by

examining the concept of a semantic void. Then, it explores a few exemplary pre-World War II

1 Lorraine Daston, ed., Biographies of Scientific Objects (Chicago: University of Chicago Press, 2000).

34 35

holistic understandings of the planet, focusing on what set these earlier ideas apart from those

that began to emerge in the 1960s and beyond. With the new tools of satellites, computer

modeling, and systems thinking scientists could now, for the first time, amass global empirical

data and use it to support world-scale theories about Earth processes. The chapter ends with an

overview of certain general trends in Earth science research in the 1970s and early 1980s,

focusing on the emergence of an institution (the National Center for Atmospheric Research), an

international scientific research program (the Global Atmospheric Research Programme), and an

individual hypothesis (James Lovelock’s Gaia hypothesis). All were part of the growing

semantic void that was created by the interdisciplinarity and global orientation facilitated by

satellites and computer modeling. Subsequent chapters will show how the ESSC’s “Earth

system” phrase filled this semantic void and provided a common term for Earth scientists to use

when referring to the Earth as a scientific object.

A SEMANTIC VOID

In “Technology: The Emergence of a Hazardous Concept,” Leo Marx observes that the meaning of the term “technology” changed in the early part of the twentieth century. “Technology” no

longer meant the field of study focused on the mechanical arts or a treatise in those arts, but

instead referred to the mechanical arts, the machines, themselves.2 Marx suggests that technology’s meaning shifted in response to broader social, scientific, and mechanical changes occurring throughout the nineteenth and early twentieth centuries that created what he calls a

“semantic void.”3 Literary and political figures struggled to find a word or phrase that

2 Leo Marx, “Technology: The Emergence of a Hazardous Concept,” Technology and Culture 51, no. 3 (Jul. 2010): 561-77. 3 Ibid, 563.

36

adequately referred to new kinds of machines and systems. Marx argues that, by the end of the

nineteenth century, increasingly large and complex systems were constructed—notably the

railroad, telegraph, and electrical systems—that required more than just hardware for them to

work properly.4 These “extras” included new financial arrangements, standards, safety

regulations, human operators, scientific and engineering knowledge, and educational institutions.

As Marx relates, observers of the new machines and systems often expressed feelings that terms

like “machines” or “mechanical arts” seemed to be inadequate descriptions.5 Beyond the

increasing prevalence of machinery and large systems, another important change occurred

regarding the concept of “progress.” It underwent a metamorphosis in the late nineteenth

century to include advances in science and the mechanical arts, not just the achievement of social

and political goals. Advances in science and the mechanical arts were no longer simply the

means for achieving social and political goals. They were now ends in themselves.6 “Machines”

became not just a measure but the measure of progress.7 These physical and conceptual changes

left in their wake a void that could not be satisfactorily described using any contemporary terms.

For writers like Henry David Thoreau and politicians like Daniel Webster, “machines” and the

“mechanical arts” no longer sufficed as descriptions of the changing machine landscape of the

late nineteenth century.

As a consequence, Marx argues that by the 1930s the term “technology” filled the void created by these broader changes. After this period, technology referred not to a field of study

4 Scholars today, following Thomas Hughes, refer to these as “large technological systems.” See: Thomas P. Hughes, “The Evolution of Large Technological Systems,” in The Social Construction of Technological Systems: New Directions in the and History of Technology, eds. Wiebe E. Bijker, et al. (Cambridge, MA: MIT Press, 1987), 51-82. 5 Marx, “Technology,” 567-72. 6 Ibid, 564-7. 7 Michael Adas, Machines as the Measure of Men: Science, Technology, and Ideologies of Western Dominance (Ithaca, NY: Cornell University Press, 1989).

37

but to the mechanical arts themselves. To make his case, Marx draws on Raymond Williams’

approach to cultural history which examines the “interdependence” or “reflexivity” between the

meanings of certain keywords and broader social and cultural changes.8 The methodological

point is that changes in the meaning of a concept often signal historically significant cultural,

social, material, and/or intellectual changes. Williams had intended to study how culture

transformed during the rise of industrial capitalism in Britain. Instead, Williams found that the

concept of culture—along with other concepts like class, industry, democracy, and art—had been

invested with new meaning by the changes introduced by industrial capitalism.9 Marx claims,

therefore, that certain keywords are, “historical markers for the periods when they acquired new,

fundamentally altered, meanings.”10 These keywords identify periods of transformation,

whether social, political, cultural, economic, scientific, technological, or some combination

thereof.

The retooling of the phrase “Earth system” circa 1986 is a historical marker in Marx’s

sense. It reflected a broader change in Earth science communities from the 1960s into the early

1980s. In this period, scientists emphasized studying the Earth as an interconnected system by

looking at large-scale interactions between the land, oceans, atmosphere, and biota. But before

1986 there was no common way to describe this new way of thinking about and investigating the

whole planet. Admittedly, this was not exactly the semantic void described by Marx. There was

not a complete absence of any phrase to describe an interconnected planet, but rather a lack of

8 See: Raymond Williams, Culture and Society, 1780-1950 (New York: Columbia University Press, 1983). In other areas of the history of science, Ian Hacking, Peter Dear, and Lorraine Daston and Peter Galison have made similar arguments about the transformation of concepts like “probability,” “experience,” and “objectivity” respectively. See: Ian Hacking, The Emergence of Probability: A Philosophical Study of Early Ideas About Probability, Induction and Statistical Inference (New York: Cambridge University Press, 1975); Peter Dear, Discipline and Experience: The Mathematical Way in the Scientific Revolution (Chicago: University of Chicago Press, 1995); Lorraine Daston and Peter Galison, Objectivity (New York: Zone Books, 2007). 9 Marx, “Technology,” 563. 10 Ibid, 575.

38

consensus about an appropriate term.11 Scientists used a number of phrases—the global system,

the climatic system, the atmosphere-ocean-ice-earth climatic system, the terrestrial ecosystem, the global ecosystem, the global environment, the global biosphere—to describe the new systemic understanding of Earth. This multiplicity only began to condense in 1986 as a result of the work of the ESSC.

Given the relatively quick adoption, lack of contestation, and later deep entrenchment of the Earth system concept in 1986 and after, the period from the 1960s to the early 1980s may usefully be characterized as a semantic void in Earth science communities. That the Earth science community was so receptive to the “Earth system” suggests that practitioners previously lacked an adequate phrase, that there was some kind of “void” to be filled. The wide adoption of the “Earth system” by scientists beginning in 1986 marks the roots of a consensus among Earth science practitioners about how to describe this new conception of an interconnected planet amenable to interdisciplinary scientific study. It marks the emergence of the Earth as a new scientific object, picked out by the “Earth system.”

EARLY HOLISTIC CONCEPTIONS OF THE EARTH

Holistic understandings of the Earth are not, of course, a twentieth-century phenomena. As Paul

Edwards states, “The idea of ‘the world’ is probably as old as language, with as many meanings and connotations as there are cultures.” Edwards suggests that “the world” probably seemed

“boundless” for much of humanity’s existence, and only became an “immense but finite globe”

11 Using Marx’s phrase “semantic void” is also not intended to maintain a strong divide between the “semantic” and the “material.” For this dissertation, “semantic” is broadly understood as interchangeable with “conceptual” (rather than more narrowly defining it as linguistic). Just like Bruno Latour’s materially instantiated “circulating reference,” concepts are always inextricably connected to material practices, instruments, concerns, actions, and the like. See: Bruno Latour, Pandora’s Hope: Essays on the Reality of Science Studies (Cambridge, MA: Harvard University Press, 1999), ch. 2.

39 in the seventeenth century. However, the ability to actually study and comprehend, “the forces that act upon it as a whole—as a system—remained for the most part beyond reach.” According to Edwards, though scientists had begun developing, “theories of world-scale processes at least as early as Copernicus, grounded empirical knowledge of geophysical features and processes remained in a rudimentary state until the Second World War.”12 Edwards provides few details about these “theories of world-scale processes” beyond mentioning Copernicus, but he might have been referring to a number of different pre-World War II thinkers. Two likely candidates include Alexander von Humbolt and Vladimir Vernadsky, each sketching sophisticated, though speculative, accounts of the holistic organization of natural entities and processes on the planet.13

The Prussian polymath (1769-1859) had a keen interest in discerning unifying or underlying patterns from empirical data.14 Susan Faye Cannon famously called his work—and the work of those following his methods—“Humboldtian Science,” characterizing it as a new and distinctive form of natural inquiry arising in the early nineteenth century that focused on quantitative measurements and their synthesis into overarching theories regarding the underlying interconnectedness of all natural entities and phenomena.15 Though

12 Paul N. Edwards, “The World in a Machine: Origins and Impacts of Early Computerized Global Systems Models.” in Systems, Experts, and Computers: The Systems Approach in Management and Engineering, World War II and After, eds. Agatha C. Hughes, Thomas P. Hughes (Cambridge, MA: MIT Press, 2000), 221. 13 Pre-WWII accounts of an interconnected planet abound. In a recent example, Sarah Dry’s Waters of the World (2019) examines how scientists have historically conceived of a holistic Earth through their studies of water in its various forms (e.g. glaciers, water vapour, precipitation, and ocean currents). See: Sarah Dry, Waters of the World: The Story of the Scientists Who Unraveled the Mysteries of Our Oceans, Atmosphere, and Ice Sheets and Made the Planet Whole (Chicago: University of Chicago Press, 2019). 14 While beyond the scope of this dissertation, Mary Louise Pratt provides a postcolonial critique of Humboldt’s travels through South America in which Humboldt viewed the and peoples he encountered in predominantly imperial terms, as resources to be used in the service of European empire-building. Aaron Sachs argues that, under close reading, Humboldt’s writings provide a foundation for a “healthy post-colonial environmentalism.” See: Mary Louise Pratt, Imperial Eyes: Travel Writing and Transculturation (New York: Routledge, 1992); Aaron Sachs, “The Ultimate “Other”: Post-Colonialism and Alexander Von Humboldt's Ecological Relationship with Nature,” History and Theory 42, no. 4 (Dec. 2003): 111. 15 Susan Faye Cannon, Science in Culture: The Early Victorian Period (New York: Science History Publications, 1978), ch. 3; Malcolm Nicolson, “Alexander von Humboldt, Humboldtian Science and the Origins of the Study of Vegetation,” History of Science 25, no. 2 (Jun. 1987): 167.

40

Humboldt emphasized the importance of collecting empirical data, the ultimate aim of his

inquiry was what he called the “physique générale,” the “universal science” that would describe

this underlying natural unity.16 Humboldt followed Immanuel Kant’s idealism, holding that the

“system of nature” was only artificially divided into classificatory schemes. The ultimate goal of

inquiry was not this artificial classification, but the development of a deeper understanding of the

underlying unity of the “system of nature.”17 Travelling throughout Europe, South and Central

America, and Siberia, Humboldt amassed as much data as he could on an immense range of

subjects (plants, animals, geological features, atmospheric conditions, and water composition)

using the most sophisticated and reliable instruments available. Humboldt believed that, with

this empirical data, he and others could ultimately “deduce” the unobservable unifying principles

of the universe. The pinnacle of Humboldt’s work was his publication of Kosmos (appearing in

five volumes between 1845 and 1862), in which he attempted to accumulate and synthesize

comprehensive data on everything in the universe, from astronomy to botany to meteorology to

anthropology. But Humboldt did not envision his Kosmos as an endpoint. He maintained that

others should continue the process of accumulation and synthesis.18

Throughout his research travels and in his writings, Humboldt sought to turn the whole

Earth (and, ultimately, the whole universe) into a single object of inquiry, into a Dastonian scientific object. The holistic science Humboldt envisaged, “can progress only by individual studies and by connecting together all the phenomena and productions on the surface of the earth. In this great chain of causes and effects, no single fact can be considered in isolation….[T]he study of nature, which is the main problem of the general physics [universal

16 Cannon, Science in Culture, 80; Nicolson, “Alexander von Humboldt,” 174. 17 Nicolson, “Alexander von Humboldt,” 171. 18 Cannon, Science in Culture, 73-110; Nicolas, “Alexander von Humboldt,” 167-94.

41

science], demands the gathering together of all the knowledge dealing with modifications of

matter.”19 Despite this aim, even Humboldt’s many resources and networks of observers could not come close to producing global datasets for the planet (let alone the universe). Using his collected observations, the most Humboldt could produce was a visual “physical tableau” depicting flora, fauna, and other physical phenomena for a regional cross section of the Andes mountain range from the Atlantic to Pacific Oceans, running through Mount Chimborazo, one of the highest mountain peaks in the Andes.20 Humboldt described this regional tableau as

containing, “almost the entirety of the research I carried out during my expedition in the

tropics.”21

By presenting these measurements and data—for plants, animals, forms of agriculture, geological structures, and atmospheric phenomena—for one region, Humboldt hoped to suggest

“unexpected analogies” that would not be readily apparent, if recognizable at all, by studying the

data in isolation.22 To facilitate the comparison of species and physical conditions with other similar climatic areas in other parts of the world like the European Alps, Humboldt included these references on the tableau.23 As historian of science Malcolm Nicolson notes, “The object

was to give, at a single glance one might say, a complete impression of a natural region—

‘régions équinoctiales’ of South America.”24 This contributed to Humboldt’s identification of

19 Alexander von Humboldt and Aimé Bonpland, Essay on the of Plants, ed. Stephen T. Jackson, trans. Sylvie Romanowski (Chicago: University of Chicago Press, 2008), 79. 20 Humboldt noted that the physical phenomena depicted in the tableau were: vegetation, animals, geological phenomena, cultivation, air temperature, limit of perpetual snow, chemical composition of the atmosphere, electrical tension, barometric pressure, decrease in gravity, intensity of the azure colour of the sky, the weakening of light as it passed through the atmosphere, the horizontal refractions of light, and the temperature of boiling water at various altitudes. See: Humboldt and Bonpland, Essay, 78. 21 Humboldt and Bonpland, Essay, 79. 22 Ibid. 23 Sylvie Romanowski, “Text of Humboldt’s Tableau physique,” in Essay on the Geography of Plants, ed. Stephen T. Jackson, trans. Sylvie Romanowski (Chicago: University of Chicago Press, 2008), 145-55. 24 Nicolson, “Alexander von Humboldt,” 178.

42

patterns of vegetation in different terrestrial locations with similar climatic conditions.25 The

important point here is that, impressive as this collection and presentation of data was, it fell well

short of providing empirical observations beyond the regional. All this map—and Humboldt’s

other visual representations of natural phenomena and entities—could do is present relationships

between geological, climatological, and biological data for sub-continental regions. Humboldt’s

ultimate vision may have been to find the structure that unified all particular, disconnected

phenomena, but his limited ability to collect and coordinate empirical data left his dream for

Kosmos unrealized.

Vladimir Vernadsky’s (1863-1945) biosphere concept may lack the grandeur of

Humboldt’s speculative universal science, but it represents another important pre-WWII holistic conception of the planet. Vernadsky was a Russian (later Soviet) mineralogist and crystallographer whose subsequent interests broadened to include , geochemistry, and biology. His research on and theorization of the biosphere concept is usually taken to presage the field that is today called biogeochemistry.26 Austrian geologist Eduard Suess first coined the

term “biosphere” in 1875 to denote the layer on the Earth’s surface where organisms live. This

layer served as an intermediary between the Earth’s crust and the air above. Suess added the

biosphere to the other “spheres” that he believed comprised the planet, like the atmosphere,

lithosphere, and .27 Historian and philosopher Jacques Grinevald credits Vernadsky with playing a crucial role in adopting and promoting this “geospheres” view of the planet.28

25 See: Nicolson, “Alexander von Humboldt”; Paul N. Edwards, A Vast Machine: Computer Models, Climate Data, and the Politics of Global Warming (Cambridge, MA: MIT Press, 2010), 30-1. 26 Alexej M. Ghilarov, “Vernadsky's Biosphere Concept: An Historical Perspective,” The Quarterly Review of Biology 70, no. 2 (Jun. 1995): 194. Scientists working in the biogeochemical tradition, particularly the Earth System Sciences Committee’s Berrien Moore, were strong supporters of conceiving of and studying the entire planet as a system analyzable in terms of feedback loops of flows of energy and matter. 27 The Oxford English Dictionary traces the roots of the term “atmosphere”—referring to a “spheroidal gaseous envelope surrounding any of the heavenly bodies”—to John Wilkins’ The Discovery of the World in a Moone

43

Vernadsky further developed Suess’ biosphere concept, most notably with his publication of The Biosphere in Russian in 1926.29 Vernadsky’s biosphere incorporated Suess’ “special cover” of the Earth’s surface that was “embraced by life,” which is to say that the biosphere was that space on Earth where organisms could be found. However, Vernadsky went further, emphasizing the importance of life on the planet, in particular its geochemical importance.

Contra many mainstream contemporary biological theories, Vernadsky maintained that organisms played an active role in creating and maintaining Earth’s conditions.30 The collective biosphere served as a “region of transformers” that converted electromagnetic radiation from the (primarily the sun) into workable chemical energy on Earth.31 The list of life’s effects on the planet was long and included what Vaclav Smil calls “obvious manifestations” like the formation of fossil fuel, carbonate, and phosphate deposits, soil creation, and the continual production of atmospheric gases like methane and oxygen that keep the atmosphere in a dynamic state of chemical disequilibrium (anticipating James Lovelock’s Gaia Hypothesis by over 40 years).32

(1638), where Wilkins wrote that, “There is an Atmo-sphæra, or an orbe of grosse vaporous aire, immediately encompassing the body of the Moone.” By the end of the nineteenth century, the OED notes that the “lithosphere” (the Earth’s crust) and “hydrosphere” (the Earth’s waters) were also in use. According to historian and philosopher Jacques Grinevald in his introduction to the first complete English language translation of Vernadsky’s The Biosphere in 1998, the biosphere became another of the Earth’s spheres, later to be joined by the “troposphere” (the lowest layer of the atmosphere) and “stratosphere” (the upper atmospheric layer above the troposphere) in 1902, the “asthenosphere” (the upper layer of the Earth’s mantle) in 1914, and the “pedosphere” (the Earth’s outermost layer composed of soil) in 1938. Though Grinevald does not mention it, the also came into usage in 1935. See: Oxford English Dictionary, accessed 16 Jul 2019, https://www.oed.com/. 28 Jacques Grinevald, “Introduction: The Invisibility of the Vernadskian Revolution,” in The Biosphere, trans. David B. Langmuir (New York: Copernicus, 1998), 23. 29 Vernadksy’s Biosphere was published in French in 1929. A complete English translation was not available until 1998. See: Ghilarov, “Vernadsky's Biosphere Concept,” 196; , et al., “Forward to the English Language Edition,” in The Biosphere, trans. David B. Langmuir (New York: Copernicus, 1998), 14. 30 Vaclav Smil, The Earth’s Biosphere: Evolution, Dynamics, and Change (Cambridge, MA: MIT Press, 2002), 4. 31 Vladimir I. Vernadsky, The Biosphere, trans. David B. Langmuir (New York: Copernicus, 1998), 43-51. 32 Smil, The Earth’s Biosphere, 4; Vernadsky, The Biosphere, 87.

44

Vernadsky recounts that one aim of The Biosphere was, “to draw the attention of

naturalists, geologists, and above all biologists to the importance of the quantitative study of the relationship between life and the chemical phenomena of the planet.”33 To better understand this relationship required, among other things, studying the cycling of certain important biogeochemical elements—namely oxygen, carbon, nitrogen, sulphur, and phosphorus—through the lands, waters, ice, and air via the life, death, and decomposition of living matter.34 Grinevald

suggests that, “Vernadsky’s Biosphere concept was part of the new geochemical point of view

that considered Earth as a dynamic energy-matter organization, a system comparable to a thermodynamic engine.” Grinevald goes on to argue that at the beginning of the twentieth century, thermodynamics (itself emerging out of industrialization engineering insights regarding energy flows and transformations), “connected with physiology, biochemistry, and (later) ecology…[and] was pivotal in the emergence of the concept of Earth as an evolving system powered by internal and external energy sources.”35

Though Vernadsky maintained the importance of studying individual organisms, the

ultimate objective was to understand the relationships of organisms and their environments at the

planetary level and how they influenced conditions on Earth. According to Vernadsky, “Because

no chemical force on Earth is more constant than living organisms taken in the aggregate, none is

more powerful in the long run.”36 Vernadsky’s world-scale theorizing fit with his intellectual

pedigree. In his history of Vernadsky’s biosphere concept, Alexej Ghilarov notes the early

influence of Humboldt and Kosmos’ universe-scale theorizing that eschewed a

33 Vernadsky, The Biosphere, 38. 34 Vernadsky, The Biosphere, 56. Textbooks on biogeochemistry (at least after the English language translation of The Biosphere in 1998) often cited Vernadsky’s ideas as precursors and contributors to this field of study. For example, see: Michael C. Jacobson, et al., Earth System Science: From Biogeochemical Cycles to Global Change (New York: Elsevier Academic Press, 2006), 6. 35 Grinevald, “Introduction,” 26. 36 Vernadsky, The Biosphere, 56.

45

compartmentalization of nature.37 Like Humboldt before him, Vernadsky claimed that all

empirical facts needed to be considered, “from the point of view of a holistic mechanism that

combines all parts of the planet in an indivisible whole. Only then will we be able to perceive

the perfect correspondence between this idea and the geological effects of life.”38 Vernadsky

insisted that he would not speculate beyond the empirical facts at his disposal, and the empirical

facts available in the 1920s could not support his Humboldtian ambitions. Vernadsky lamented

that his theories had been, “limited to the scant number of precise observations and experiments

at my disposal.” Vernadsky saw The Biosphere as a means to urge scientists to undertake data collection endeavours: “A great number of quantitatively expressed empirical facts need to be collected as rapidly as possible.”39 This was required before empirically-supported global

scientific theories of the Earth could be developed. Vernadsky wanted to construct a

Humboldtian map of an interconnected planet where life played a strong role in creating and

maintaining planetary conditions. That hope and vision was thwarted by a lack of empirical

data.

THE MID-TWENTIETH CENTURY SEMANTIC VOID

Humboldt and Vernadsky’s world-scale theories suggest that, though they were planet-

encompassing (or more), they were based primarily on speculative reasoning on the basis of a

few examples rather than global empirical data. Both envisaged a unified global model but they

were frustrated by the inability to collect empirical data on a global scale. What the above shows

is that, though there were attempts to theorize globally, there was not, prior to the mid-twentieth

37 Ghilarov, “Vernadsky's Biosphere Concept,” 194. 38 Vernadsky, The Biosphere, 40 [emphasis original]. 39 Ibid, 38.

46

century, the data collection capacity to actually do so. For the whole Earth to become an object

of scientific inquiry it was not enough for natural historians or philosophers or scientists to

develop grand theories about the planet.40 They needed to collect and analyze data on a global

scale to solidify these theories. This was not practically possible before the mid-twentieth

century. According to Daston in Biographies of Scientific Objects (2000), a scientific object is

not merely something of interest to scientists. Such a definition would make almost anything a

scientific object. Unlike stable, familiar, obvious, unavoidable, quotidian objects—the tables and

books and walls and laptops and clothing of our everyday experience—scientific objects are something more rare, something “elusive and hard-won.”41 For Daston, a scientific object is

something for which there are techniques that can take previously unknown or ignored or

dispersed phenomena and coalesce them into a unified representation that becomes accepted

within a discipline or disciplines. This representation can then, according to Daston, “be

observed and manipulated,” and “is capable of theoretical ramifications and empirical

surprises.”42 An object may be real in a commonsense way before becoming an object of

scientific inquiry, but when it becomes a scientific object it acquires a new salience in a domain

of inquiry. Thus, “scientific objects attain their heightened ontological status by producing

results, implications, surprises, connections, manipulations, explanations, applications.”43 The

whole Earth, as theorized by Humboldt and Vernadsky, did not have this heightened ontological

status, but rather remained predominantly conjectural.

This situation began changing in the 1960s with the emergence of new technologies and

techniques. In his examination of the “global knowledge infrastructure” of climate science, Paul

40 Daston, ed., Biographies of Scientific Objects. 41 Lorraine Daston, “Introduction: The Coming into Being of Scientific Objects,” in Biographies of Scientific Objects, ed. Lorraine Daston (Chicago: University of Chicago Press, 2000), 2. 42 Ibid, 5. 43 Ibid, 10.

47

Edwards identifies two major technological components that both facilitated a new conception of

the planet and provided the tools to actually study it as a singular Earth system. These were the

tools that allowed “the world” to become a scientific object (though Edwards does not use that

specific phrase): Earth observing satellites and global computer models.44 As Edwards notes,

before satellite data and global computer modeling, there were no tools available for scientists to

make global knowledge about the planet with an empirical basis.45 With these tools came the

belief that scientists in the separate Earth science disciplines could and should engage in greater

collaboration to develop a more comprehensive understanding of the whole planet. The arrival

of satellite data collection and computer modelling in the 1960s meant that the entire Earth

could, for the first time, become an object of scientific inquiry.

Earth Observing Satellites

The history of each of these new tools merits an (admittedly brief) overview. To begin with

satellites, the first artificial Earth orbiting satellite (Sputnik 1) was launched by the USSR on

October 4, 1957 as part of the International Geophysical Year (IGY), a major international endeavour involving thousands of scientists from 67 countries to conduct coordinated

observations of many of the Earth’s geophysical properties.46 Edwards identifies the IGY as

pivotal in the global extension of observational networks—what he calls “infrastructural

44 Paul N. Edwards, “Representing the Global Atmosphere: Computer Models, Data, and Knowledge About Climate Change,” in Changing the Atmosphere: Expert Knowledge and Environmental Governance, eds. Clark A. Miller, Paul N. Edwards (Cambridge, MA: MIT Press, 2001), 31-66; Edwards, A Vast Machine; Edwards, “The World in a Machine,” 221-53. 45 A number of Edwards’ articles and books have made this argument. See: Edwards, A Vast Machine; Edwards, “The World in a Machine”; Edwards, “Representing the Global Atmosphere.” 46 Roger D. Launius, James Rodger Fleming, and David H. DeVorkin, eds., Globalizing Polar Science: Reconsidering the International Polar and Geophysical Years (New York: Palgrave Macmillan, 2010).

48 globalism”—and the emergence of the idea to study the planet as a “single .”47

The earliest Earth orbiting satellites fit into three broad categories: military spy satellites (e.g. the

US’ Corona Program), communications satellites (e.g. the AT&T-operated Telstar), and scientific satellites intended to study the Earth, simply called Earth observing satellites.48

Some of the earliest Earth observing satellites were launched to collect meteorological data. This included the TIROS series of weather satellites that NASA took over from the US

Army Ballistic Missile Agency in 1959.49 Conway reports that, in its short 76-day life span

( with the satellite was lost on June 15, 1960), TIROS 1 transmitted 22,952 usable images to ground stations on Earth.50 These cloud-cover images were immediately useful for

47 Paul N. Edwards, “Knowledge Infrastructures for the Anthropocene,” The Anthropocene Review 4, no. 1 (2017): 36-7; Edwards, A Vast Machine, 204. In addition to Edwards’ work, which looks at satellites as they contributed to climate science, there are two other book-length histories of Earth observing satellites. Pamela Mack’s Viewing the Earth (1990) provides a history of NASA’s Landsat satellite program, from its conception in the 1960s up to the launch of Landsat 4 (March 1, 1984) and also explores NASA’s attempts to generate user groups for the collected data. Erik Conway’s Atmospheric Science at NASA (2008) also focuses on NASA, but this time examining not just NASA’s atmospheric science research, but also its broader Earth and environmental science efforts. A recent book- length treatment by Richard Leshner and Thor Hogan examines NASA's use of Earth observing satellites to study the planet. There are also two dissertations that outline the observing satellites at NASA by Eric Goldstein and Richard Leshner. See: Pamela E. Mack, Viewing the Earth: The Social Construction of the Landsat Satellite System (Cambridge, MA: MIT Press, 1990); Erik M. Conway, Atmospheric Science at NASA: A History (Baltimore: Johns Hopkins University Press, 2008); Edward S. Goldstein, “NASA’s Earth Science Program: The Bureaucratic Struggles of the Space Agency’s Mission to Planet Earth” (PhD dissertation, George Washington University, 2007); Richard B. Leshner, “The Evolution of the NASA Earth Observing System: A Case Study in Policy and Project Formulation” (PhD dissertation, George Washington University, 2007); Richard B. Leshner and Thor Hogan, The View From Space: NASA's Evolving Struggle to Understand Our Home Planet (Lawrence, KS: University Press of Kansas, 2019). 48 For more information on the Corona Program, see: John Cloud, “Imagining the World in a Barrel: CORONA and the Clandestine Convergence of the Earth Sciences,” Social Studies of Science 31, no. 2 (Apr. 2001): 231–51. For more on telecommunications satellites, see: Hugh Richard Slotten, “Satellite Communications, Globalization, and the Cold War,” Technology and Culture 43, no. 2 (Apr. 2002): 315-50. 49 Confusingly, Mack lists two different years for when the transfer of Tiros to NASA occurred. In Viewing the Earth, Mack lists the year as being 1959. In “Observing the Earth from Space,” Mack and Ray Williamson state the year as being 1958. See: Mack, Viewing the Earth, 20; Pamela E. Mack and Ray A. Williamson, “Observing the Earth from Space,” in Exploring the Unknown: Selected Documents in the History of the US Civil Space Program, Vol. III: Using Space, ed. John M. Logsdon (Washington, DC: NASA, [1998]), 158. Conway helps clear up the confusion by noting that President Dwight Eisenhower decided the day before the Space Act legislation was signed (in 1958) to transfer control of all meteorological satellites to NASA, which would be arranged for 13 April 1959. See: Conway, Atmospheric Science at NASA, 28. 50 Conway, Atmospheric Science at NASA, 29. Oddly, NASA itself reports two slightly different lifespans for TIROS 1. On their “NASA Science Missions” website, NASA reports that the satellite was operational for 78 days. However, on NASA’s National Space Science Data Center website, NASA states that on 15 June 1960 an “electrical

49 meteorological purposes, revealing large-scale trends and weather patterns not easily discernible from ground observations.51 The TIROS program led to a series of weather satellites developed and tested by NASA, including the Nimbus program, consisting of seven satellites carrying a variety of electromagnetic sensors.52 To provide just one example of the data collecting potential of these satellites, Nimbus 1—with instruments to collect cloud cover images and nighttime radiative data—launched on August 28, 1964 and lasted only 26 days, but in that time collected more than 27,000 images.53 NASA developed these meteorological satellites in parallel with its Landsat series, intended to collect various kinds of data on the Earth’s land surfaces. NASA launched Landsat 1 on July 23, 1972 mounted with two remote sensors: a return beam vidicon (a type of television camera with a very high resolution), and a multispectral scanner (which collected radiometric data from the Earth in four spectral bands).54 With almost every new satellite in the Landsat program, the mounted instruments became more sophisticated and collected increasing quantities of data, as new electromagnetic bands (e.g. to detect thermal infrared radiation) were added to the multispectral scanner and a thematic mapper.55

power failure prevented further useful TV transmission.” This would give the satellite an operational lifespan of 76 days. See: “TIROS, Television Infrared Observation Satellite Program,” NASA, accessed 10 June 2015, http://science1.nasa.gov/missions/tiros/; “TIROS 1,” NASA, accessed 10 June 2015, http://nssdc.gsfc.nasa.gov/nmc/spacecraftDisplay.do?id=1960-002B. 51 Mack, Viewing the Earth, 20; Conway, Atmospheric Science at NASA, 29; Edwards, A Vast Machine, 221. 52 Conway, Atmospheric Science at NASA, 38-63. 53 Nimbus 1 carried three instruments: an Advanced Vidicon Camera System (AVCS), the Automatic Picture Transmission (APT), and a High Resolution Infrared Radiometer (HRIR). See: Conway, Atmospheric Science at NASA, 41-3. 54 The return beam vidicon used a shutter to expose a plate of semiconducting material to light. An electron beam then scanned this plate, recording a signal when the beam bounced back from a spot on the plate electrically charged by its exposure to light. This data was then radioed to a data collection center on Earth. See: Mack, Viewing the Earth, 68-70. The multispectral scanner used a mirror to reflect light from the Earth onto an array of photoelectric detectors, which produced an electric signal that varied based on the intensity of the reflected light. See: Mack, Viewing the Earth, 70-73. 55 Subsequent satellites were launched in 1975, 1978, 1982, 1984, 1993, 1999, and 2013. The satellite launched in 1993, Landsat 6, was the only one that failed to achieve orbit. The most recent launch was of Landsat 8 on 11

50

The trend of launching more satellites with ever more sophisticated remote sensing

instruments that collected ever more amounts of data continued into the 1980s for satellites

observing conditions in the atmosphere and on land, as well as on the oceans. In Conway’s

examination of NASA’s Seasat, he notes that, though Seasat failed after 106 days in 1978, if it

had functioned for its intended three year lifespan the satellite’s five instruments would have

produced 6,000 data tapes per year, containing tens of thousands of measurements.56 This is orders of magnitude more data than could have been collected from shipboard instruments. This growth in data collection capacity prompted the Scripps Institute of Oceanography’s director

William Nierenberg to remark that, “[the] amounts and kinds of data that will become available to the oceanographic community from satellites such as Seasat-A are completely foreign to the experience of the existing academic oceanographic community.”57 This was a revolutionary

period for those in the Earth sciences, with vastly more data becoming ever more readily

available for ever more components of the planet. As Conway observes, the 1960s and 1970s

was a period when institutions like NASA, the National Oceanic and Atmospheric

Administration (NOAA), and the National Center for Atmospheric Research (NCAR) developed

and tested new remote sensing instruments and made them ready to fruitfully contribute to

scientific endeavours. Conway writes: “these new measurement technologies seemed capable of

February 2013, with Landsat 9 being slated for launch in 2023. See, Mack, Viewing the Earth, 198-200; “Landsat Science: History,” NASA, last updated 10 June 2015, http://landsat.gsfc.nasa.gov/?page_id=2281; “The Thematic Mapper,” NASA, last updated 10 June 2015, http://landsat.gsfc.nasa.gov/?p=3229. 56 Seasat’s five instruments were: a microwave altimeter (to measure sea surface height), a microwave imaging radiometer (to measure sea surface temperatures under all conditions), a synthetic aperture radar (to measure profiles of sea ice, sea state, and potentially internal waves), a scatterometer (to measure surface wind speeds), and another imager operating in the visible and infrared spectrums (to collect sea surface temperatures in cloud-free areas). See: Erik M. Conway, “Drowning in Data: Satellite Oceanography and Information Overload in the Earth sciences,” Historical Studies in the Physical and Biological Sciences 37, no. 1 (2006): 135-6, 140. 57 Quoted in: Conway, “Drowning in Data,” 139-40 [emphasis added].

51 producing global meteorological datasets, permitting the construction and validation of global numerical prediction models.”58

Computer Modeling

Earth observing satellites were collecting massive amounts of data that, for the first time, allowed practitioners to seriously believe that the datasets were global in scope. However, without the analytical capacities provided by digital computer modeling, it is highly unlikely that the new surfeit of satellite data would have been of much use. Edwards argues that global computer models drove the development of a global observation system dependent on Earth observing satellites since these models created, “an epistemological framework in which gathering global information became necessary and made sense.” It was the modeling efforts of those working in the Earth sciences that Edwards claims, “provided the rationale for the creation of global data networks.”59 Global models required global data for their construction, calibration, and validation. Without these models, there would have been no reason to collect satellite data since there would have been no way to analyze it, and without analysis there would be no way to use the data to produce scientific knowledge.

The kinds of satellite data collected varied according to the needs of computer models for meteorology, oceanography, and other fields. For instance, the images of cloud cover produced by the TIROS satellites were useful for qualitatively identifying certain atmospheric conditions like large low pressure areas, but as Conway notes, visual information could not be used in numerical weather models that described the atmosphere’s motion using quantitative

58 Conway, Atmospheric Science at NASA, 41. 59 Edwards, “The World in a Machine,” 244, 246.

52

measurements of atmospheric temperature, pressure, wind, and the like.60 Similarly, data collected during the IGY was more comprehensive than anything previously collected, but the data were stored in analog format on microfilm, a format unusable in numerical models running on digital computers.61 Numerical models represented a transition away from the pictorial

mapping of sub-continental regions that incorporated collected data into these images. Maps,

like Humboldt’s Mount Chimborazo profile or a weather map depicting warm and cold fronts,

could be useful for presenting, at a glance, relationships between different kinds of

measurements for a single region. However, this kind of map could only present a static slice of

measurements at a single time. It was not capable of presenting a dynamic picture of conditions

on the planet using a time series of measurements. It was not capable of predicting future states

of the planet using present data and physical theories about the flows of gases and liquids.

Edwards provides an extensive historical study of the development of computer models

in the Earth sciences, focusing primarily on climate science and weather forecasting. Other

Earth and environmental science histories—notably those by Erik Conway, Kristine Harper,

James Fleming, Joel Hagen and Sharon Kingsland—cover features of the history of computer models as part of broader histories of meteorology, atmospheric science, oceanography, and ecology.62 There is, as yet, no scholarship focused solely on the history of computer modeling in

the Earth and environmental sciences. The first numerical models of Earth processes described

meteorological phenomena mathematically in an effort to improve weather forecasts. Beginning

60 Conway, Atmospheric Science at NASA, 40. 61 Elena Aronova, “Geophysical Datascapes of the Cold War: Politics and Practices of the World Data Centers in the 1950s and 1960s,” Osiris 32 (2017): 307-27. 62 Conway, Atmospheric Science at NASA; Kristine C. Harper, Weather by the Numbers: The Genesis of Modern Meteorology (Cambridge, MA: MIT Press, 2008); James Rodger Fleming, Inventing Atmospheric Science: Bjerknes, Rossby, Wexler, and the Foundations of Modern Meteorology (Cambridge, MA: MIT Press, 2016); Joel B. Hagen, An Entangled Bank: The Origins of (New Brunswick, NJ: Rutgers University Press, 1992); Sharon E. Kingsland, The Evolution of American Ecology, 1890-2000 (Baltimore, MD: Johns Hopkins University Press, 2005).

53 in the 1910s, Lewis Fry Richardson engaged in an early attempt to predict weather using numerical calculations.63 Taking a set of observational data for central Europe from a single day and using the complete hydrodynamical equations, Richardson took 11 years to construct a six- hour forecast that was of doubtful fidelity and, due to its lack of timeliness, was clearly not usable as an actual weather forecast.64 Richardson’s basic idea developed into what is now called Numerical Weather Prediction (NWP).65 What has changed is the way that the calculations have been carried out. Richardson did his calculations by hand. These calculations could be performed by machines after the development of electronic digital computers during and after World War II.

A meteorological forecast was one of the first calculations completed with the US’s first electronic digital computer, the Electronic Numerical Integrator and Computer, or ENIAC. In

March and April 1950, computer pioneer John von Neumann—assisted by theoretical meteorologists like Carl-Gustav Rossby and the “ENIAC girls”—employed the ENIAC to simulate Earth processes relating to the atmosphere using a highly simplified mathematical model of the physical motion of the atmosphere.66 It took 33 days (around 800 computer hours) of around-the-clock work to produce two 12 hour and four 24 hour retrospective weather

63 Fleming notes that Felix Exner and Vilhelm Bjerknes conducted more limited numerical weather forecasting calculations before Richardson. See: Fleming, Inventing Atmospheric Science, 65. 64 Fleming notes that Richardson calculated an unrealistic 40-millibar pressure rise. See: Fleming, Inventing Atmospheric Science, 64. 65 Edwards provides an excellent description of NWP that uses mathematical modeling and digital computers: “NWP models, then as now, work by breaking up the atmosphere into a set of ‘grid boxes,’ tens to hundreds of kilometers square and hundreds to thousands of meters deep. Within each grid box, conditions such as temperature, humidity, and pressure are assumed homogeneous. The models simulate what happens to the air mass in each grid box on a period ‘time step’ (today, typically about twenty to thirty minutes; in early models, sometimes as much as three hours).” See: Edwards, “The World in a Machine,” 225. 66 The “ENIAC girls” were six women—Kathleen McNulty, Frances Bilas, Betty Jean Jennings, Betty Holberton, Ruth Lichterman, and Marlyn Wescoff—recruited to help wire the ENIAC to perform calculations. They are today considered the first computer programmers. See: Jennifer Light, “When Women Were Computers,” Technology and Culture 40, no. 3 (Jul. 1999): 455-83. Edwards notes that, in order to produce a model simple enough to be calculable by the ENIAC in a relatively short period of time, the three equations of motion were reduced to one, with pressure being the only dependent variable. Other simplifying assumptions for the sake of expediency were made as well. See: Edwards, A Vast Machine, 119-20.

54

“forecasts.”67 Despite the long time to produce those calculations and the “far from perfect” results, the proof of concept had been achieved.68 Weather forecasting using NWP calculation

techniques could be conducted using digital computers, with faster computers promising faster

calculations that might eventually amount to the capability for real-time weather prediction. The

first routinized NWP used by national weather services on a hemispheric scale began in Sweden

in 1954 and the US in 1955.69

Following the ENIAC’s “success,” models simulating global atmospheric motion

(General Circulation Models, GCMs) were developed to simulate, first, the world’s weather and, later, the world’s climate.70 By 1953, von Neumann and other meteorologists (notably the US

Weather Bureau’s Harry Wexler) proposed government-sponsored research into numerical

methods to model the planet’s general atmospheric circulation. This resulted in a number of

multi-layer continental, hemispheric, and eventually global GCMs being developed around the

US.71 These models required the computation of changes in thousands of so-called grid boxes

around 20,000 times to simulate the world’s climate for a single year. Between 1970 and 1980,

the computer time required to run a typical declined tenfold. In 1971, it was estimated that a climate model required roughly 120 hours of computer time to model global

67 This model was a two-dimensional 15 x 18 grid that included 270 grid points at intervals of 736km, which covered North America and many of the surrounding ocean waters. The model incorporated a three-hour time step conducted at the level of 500 millibars, roughly six kilometers above the Earth’s surface. See: Edwards, A Vast Machine, 120. 68 Edwards, “The World in a Machine,” 225. 69 Ibid. 70 GCMs are simply global versions of NWP models. When these models are used for weather forecasting, they are initiated using observational data and only provide simulations for a period of up to two weeks. Climate models are usually not initiated with observational data but begin instead with physical constants (e.g. incoming solar radiation, the speed of the Earth’s rotation, the Earth’s wobble on its axis, and atmospheric chemical composition), and run until they reach a stable, or equilibrium state, which is the model’s “climate” that can then be compared with observational data. See: Edwards, “The World in a Machine,” 231-2. 71 These US groups included: the Geophysical Fluid Dynamics Laboratory (GFDL), today located at Princeton University; a group at UCLA; a group at the Lawrence Livermore National Laboratory; and a group at NCAR. See: Edwards, “The World in a Machine,” 228-9, 231-2.

55

climate for one year. Supercomputers had, by the early 1980s, reduced this time to roughly 12

hours per climate year modeled.72

Computer modeling capabilities in other Earth science disciplines grew during this period

as well, though no scholarship has focused specifically on these efforts. Work by Edwards and

Conway that focuses primarily on atmospheric science provides a brief insight into the modeling

work undertaken by oceanographers. This research lacuna may be explained by the fact that,

according to Conway, ocean modeling was only “in its infancy” when compared to the work

being done in atmospheric science and that “substantial” oceanographic modeling did not begin until the 1980s.73 Modeling work in other fields was either just as nascent as in oceanography,

or in even more elemental stages. A report prepared by members of the Earth System Sciences

Committee’s Earth Systems Modeling Working Group (ESMWG) in 1984 nevertheless

described the state of modeling in optimistic terms:

The capabilities of dynamical, short-term models of the atmospheric and oceanic circulations are well known. Models of forest ecology that allow for evolution of the distribution of species as functions of climatological parameters are under active development. Carbon cycle models, accounting in detail for oceanic and biological processes, are available. Upper atmosphere chemistry models are progressing, and being linked to dynamical models…. Ocean ecology modelers are attacking problems related to ocean dynamics and biological processes. Interfacial processes involving evapotranspiration and surface runoff are being modeled in detail. Questions associated with dynamics and thermodynamics of crustal motion and plate tectonics are being attacked from first principles and, although with other subsystems are not yet clear, are susceptible to modeling.74

72 Despite these computational improvements, GCMs still required 1,200 hours, or 50 continuous days, of supercomputer time. Even today’s more complex GCMs still require hundreds of supercomputer hours to produce climate models. See: Edwards, “The World in a Machine,” 231. 73 Conway, Atmospheric Science at NASA, 249. 74 Earth System Sciences Committee (ESSC), Working Group on Earth System Modeling, First Meeting, Washington, D.C., October 1984, ESSC/Bretherton Committee File, Box 18042, NASA HRC, Washington, DC.

56

As the ESMWG reported, models either existed, were under development, or could be

developed, in many of the major Earth science disciplines in the mid-1980s.

Edwards describes model development as proceeding in three major ways: first, the

development of improved equations for representing physical processes and their integration into

the models’ algorithms; second, the reduction of the size of model grid boxes, to achieve greater

granularity, and; lastly, the expansion of models to include more physical processes. Edwards

lists some of the additions to climate model physics: “radiative transfer, cloud formation, ocean

circulation, albedo, sulfate emissions, and particulate aerosols.” In the 1980s, items such as sea

ice, snow, vegetation, and land use were added to climate models to create what Edwards calls

“Earth system models.” However, Edwards notes that it was not until the mid-1990s that

modelers began integrating the entire carbon cycle into these models. Nevertheless, the trend

throughout the 1980s and beyond was towards increasingly comprehensive “Earth system models” that coupled ever more components of the Earth: the atmosphere, oceans, land surfaces, cryosphere (sea ice, glaciers, snow cover), hydrosphere (lakes and rivers, evaporations, precipitation), and vegetation.75

Systems Thinking and

The computational models developed in the Earth sciences after 1960 made use of a “systems thinking” or a “systems approach” to analysis. Literary scholar Clifford Siskin begins his study of how “system” shaped modern knowledge with a question from the physicist Leonard

Susskind: “What is a system?” According to Susskind (and Siskin), this “primitive question” is unanswerable. Rather than trying to define “system,” Siskin suggests that “system” should be

75 Edwards, A Vast Machine, 146, 418.

57

studied in its different contexts, in an effort to understand how “system” has been used and how

this use has changed over time.76 “System,” like all concepts, has a history. Though the word

“system” has ancient roots, Siskin points to the work of Galileo in the seventeenth century as the time when “system” took on a modern meaning. With Galileo’s identification of moons orbiting

Jupiter and his promotion of Copernicus’ heliocentrism, the universe was no longer a singular system. The universe, after Galileo, could be considered a system comprised of other systems.

In Galileo’s view, the Earth (with the other planets) orbited the sun as part of a system, but the moon orbited the Earth in another system, with Jupiter’s moons orbiting it in yet another system, and no doubt there were other systems in the universe awaiting identification. After Galileo, there could be what Siskin calls a “proliferation of systems.”77 A system could, ever after,

contain other systems or be contained within larger systems. This made the new concept of

system scalable, applicable to any and all sizes of organizations of entities.78 Siskin notes a

dramatic increase in the number of uses of “system” in titles published beginning in the

seventeenth and eighteenth centuries. In the sciences, authors frequently described their

astronomical and physical texts as “systems,” beginning with Galileo’s Dialogue Concerning the

Two Chief World Systems (1632) and Isaac Newton’s third book of Philosophiæ Naturalis

Principia Mathematica (1687), “De mundi systemate” (On the system of the world).79

Whatever the exact roots, scientists adopted the notion of “system” into their research in

more precise ways by the early twentieth century, as a specific tool for focusing scientific

analysis and managing complexity. The discipline of ecology provides an early example. In

1934, ecologist Arthur Tansley famously promoted the use of the term “ecosystem” to refer to

76 Clifford Siskin, System: The Shaping of Modern Knowledge (Cambridge, MA: MIT Press, 2016), 1-2. 77 Ibid, ch. 1. 78 Siskin describes a system as a “scalable technology” that can be applied to all scales, large or small, which greatly enhances its conceptual utility. See: Siskin, System, 7, 29, 36. 79 Siskin, System, 2, 18.

58 the ecological area under study (rather than using misleading “organism” metaphors that risked being taken literally). Tansley drew on philosopher and mathematician Hyman Levy’s description of scientific practices in The Universe of Science (1932). Levy presented a contingent and constructed view of science where, he argued, scientists studied those aspects of,

“the changing world that are amenable to scientific analysis,” where the kinds of questions scientists asked affected the kinds of answers produced.80 Tansley described Levy’s understanding of the method of science as one where scientists,

isolate systems mentally for the purposes of study, so that the series of isolates we make become the actual objects of our study, whether the isolate be a solar system, a planet, a climatic region, a plant or animal community, an individual organism, an organic molecule or an atom. Actually the systems we isolate mentally are not only included as parts of larger ones, but they also overlap, interlock and interact with one another. The isolation is partly artificial, but is the only possible way in which we can proceed.81

The isolation of a system may be an artifice or convention, but it was, for Tansley, a practical necessity, a way of grappling with entities and phenomena of such complexity that they could not be managed in any other way. It is a practical technique for narrowing scientific focus, for slicing off only a small and workable portion of the rich and vast abundance of the universe.

Even before it became a more refined, mathematical technique, this early “systems approach” proved useful for conceptually mapping a complex world.

In the twentieth century, some scientists attempted to construct more mathematically rigorous modes of analysis that made use of the identification and analysis of systems in the world. Two notable early examples are general and cybernetics. There is limited work on general systems theory written by historians, though many current and former

80 H. Levy, The Universe of Science (London: Watts & Co., 1932), x. 81 A.G. Tansley, “The Use and Abuse of Vegetational Concepts and Terms,” Ecology 16, no. 3 (Jul. 1935): 300 [original emphasis].

59 practitioners have offered personal histories. ’s The Science of Synthesis

(2003) provides one of the few sources, and even it is more philosophical than historical.82 The

Austrian biologist (1901-1972) was one of the earliest and most prominent promoters of what he called “general system theory” (“system” was later pluralized by practitioners). In a number of publications later collected into one volume, General System

Theory (1968), von Bertalanffy explained that this theory attempted to develop the general principles that could be applied to all systems. These principles would be germane to all systems in the natural and human sciences, and possibly to systems in other fields like history, linguistics, philosophy, and education.83

To note just one example, von Bertalanffy argued that the “exponential law or law of compound interest” could be applied (with a negative integer) to a wide variety of phenomena, including: “the decay of radium, the monomolecular reaction, the killing of bacteria by light or disinfectants, the loss of body substance in a starving animal, and to the decrease of a population where the death rate is higher than the birth rate.” Using a positive integer, this law could be further applied, “to the individual growth of certain micro-organisms, the unlimited Malthusian growth of bacterial, animal, or human populations, the growth curve of human knowledge (as measured by the number of pages devoted to scientific discoveries in a textbook on the history of science), and the number of publications on Drosophila.” According to von Bertalanffy, “The entities concerned—atoms, molecules, bacteria, animals, human beings, or books—are widely different, and so are the causal mechanisms involved. Nevertheless, the mathematical law is the

82 Debora Hammond, The Science of Synthesis: Exploring the Social Implications of General Systems Theory (Boulder, CO: University Press of Colorado, 2003). 83 Ludwig von Bertalanffy, General System Theory: Foundations, Development, Applications (New York: George Braziller, 1968).

60

same.”84 For von Bertalanffy, there were structural or formal similarities between different systems that could be identified and then used in subsequent analyses. Von Bertalanffy’s work influenced what came to be called general systems theory, and was adopted by a number of practitioners in a variety of disciplines, including physiology, biology, , psychology, and anthropology.85

Similarly, practitioners in the field that came to be identified as cybernetics also

developed general principles applicable to different systems, be they animals (including human

animals) or machines. A number of scholars have examined the history of cybernetics from a

variety of perspectives that range from more engineering oriented (David Mindell) to literary

theories of posthumanism (N. Katherine Hayles) to psychology and psychiatry (Andrew

Pickering) to the adoption of cybernetic ideas by democratic socialist regimes (Eden Medina).86

Recent work by Ronald Kline examines the rise and fall of the field of cybernetics, arguing that

the eventual “triumph” of information theory over cybernetics by the 1970s resulted more from

cybernetics’ unfortunate association with pseudoscientific ideas like L. Ron Hubbard’s

Scientology than from a rejection of cybernetic principles.87 coined the term

“cybernetics” in his seminal publication Cybernetics: Or Control and Communication in the

Animal and the Machine (1948). This gave a name to a nascent line of research—similar to von

Bertalanffy’s general system theory—that identified control and communication in all kinds of

84 Ludwig von Bertalanffy, “An Outline of General System Theory,” The British Journal for the Philosophy of Science 1, no. 2 (Aug. 1950): 136. 85 Hammond’s list of notable general systems theorists includes: economist Kenneth Boulding, the psychologist , the neurophysiologist Ralph Gerard, biologist , and anthropologist . See: Hammond, The Science of Synthesis. 86 David Mindell, Between Human and Machine: Feedback, Control, and Computing Before Cybernetics (Baltimore, MD: Johns Hopkins University Press, 2002); N. Katherine Hayles, How We Became Posthuman: Virtual Bodies in Cybernetics, Literature, and Informatics (Chicago: University of Chicago Press, 1999); Andrew Pickering, The Cybernetic Brain: Sketches of Another Future (Chicago: University of Chicago Press, 2010); Eden Medina, Cybernetic Revolutionaries: Technology and Politics in Allende’s Chile (Cambridge, MA: MIT Press, 2011). 87 Ronald R. Kline, The Cybernetics Moment: Or Why We Call Our Age the Information Age (Baltimore, MD: Johns Hopkins Press, 2015).

61 substrates, technological or biological, as similar systems that could be analyzed in terms of feedback loops, information processing and exchange, homeostatic/steady states, and goal- oriented (teleological) action.88

Cybernetics was a tool for understanding and/or constructing complex organizations of entities using similar techniques to analyze seemingly different phenomena. Methods of simplifications were, according to Wiener, the only way to grapple with complexity: “One thing that we cannot do is to take the full complexity of the world without simplification of methods.

It is simply too complicated for us to grasp.”89 Wiener (and others) called for the use of mathematical techniques like statistical analysis, random functions, ergodic theory, harmonic analysis, mathematical modeling, and computer simulations to build or model systems.90 A thermometer, a gyrocompass, automatic gun fire, and a human reaching for a pen could, for instance, all be analyzed using these common techniques. While Wiener remained skeptical about the applicability of cybernetics beyond the realms of mathematics, engineering, and physiology, other attendees of the Macy Conferences—a series of meetings held from 1946 to

1953 to debate these ideas and foster interdisciplinary collaboration—were more confident that these ideas could be used in the social sciences.91 Attendees like and Margaret

Mead sought to apply these principles to anthropology, while Kurt Lewin saw important applications in the field of psychology.92 Another attendee, G. Evelyn Hutchinson, was an early promoter of cybernetics in ecology. While Vernadsky’s biosphere concept and theories of the flows of elements throughout the planet provided a conceptual grounding for Hutchinson’s

88 Norbert Wiener, Cybernetics: Or Control and Communication in the Animal and the Machine, second edition (Cambridge, MA: MIT Press, 1961). 89 Quoted in: Kline, The Cybernetics Moment, 49. 90 Wiener, Cybernetics. 91 Kline, The Cybernetics Moment, 1. 92 Steve Heims, The Cybernetic Group (Cambridge, MA: MIT Press, 1991).

62

ecosystem modeling, it was cybernetics that provided him with the mathematical tools by which

this could be achieved.93 Hutchinson would later train ecologist Howard Odum who, along with

his brother Eugene, were proponents of cybernetic and systems approaches in ecology.94

Though the word “cybernetics” would eventually fall out of use by the 1970s, cybernetic

and systems modes of analysis were enormously influential in the decades after World War II.

Thomas Hughes describes this in both an edited volume (Systems, Experts, and Computers,

2000) and as part of an examination of the use of systems approaches in four major technological

projects after World War II (Rescuing Prometheus, 2000).95 Hughes identifies three major

systems approaches that emerged from this period: , , and

. Operations research referred to the use of quantitative techniques to analyze

already deployed military and industrial systems to improve performance or results. Systems

engineering was a style of management for large complex systems comprised of many specialist

components (for instance, an intercontinental ballistic missile) that provided a high-level

coordination of constituent parts while at the same time giving experts the latitude to develop

their specific components autonomously. Systems analysis evaluated different proposed projects, often using computer models, to determine the optimal projects to achieve specific

93 Hutchinson was introduced to Vernadsky’s ideas via Vernadsky’s son George. Both taught at Yale in the 1920s and George Vernadsky translated his father’s work into English for Hutchinson. See: Smil, The Earth’s Biosphere, 11. 94 Nancy G. Slack, G. Evelyn Hutchinson and the Invention of Modern Ecology (New Haven, CT: Yale University Press, 2010). Joel Hagen argues that the Odums developed cybernetic notions of with feedback loops— what they called metabolic flows—and (self-regulation). See: Joel B. Hagen, “Eugene Odum and the Homeostatic Ecosystem: The Resilience of an Idea,” in Traditions of Systems Theory: Major Figures and Contemporary Developments, ed. Darrell P. Arnold (New York: Routledge, 2014), 179-96. 95 Agatha C. Hughes and Thomas P. Hughes, eds., Systems, Experts, and Computers: The Systems Approach in Management and Engineering, World War II and After (Cambridge, MA: MIT Press, 2000). The four projects Hughes examines in Rescuing Prometheus are: the Semi Automatic Ground Environment (SAGE), the Atlas intercontinental ballistic missile project, Boston's Central Artery/Tunnel Project, and ARPANET. See: Thomas P. Hughes, Rescuing Prometheus: Four Monumental Projects That Changed the Modern World (New York: Vintage Books, 2000).

63

objectives. All of these approaches relied on the quantitative techniques developed by general

system theory and cybernetics.96

According to Hughes, these approaches originated in military and defence planning and

development, but practitioners expanded their application during President Lyndon Johnson’s

Great Society reforms of the mid-1960s in an effort to solve social issues targeted by these

reforms (and to profit from the increased government funding available in these areas).97

Jennifer Light elaborates on this expansion of systems approaches into non-defence realms in

From Warfare to Welfare (2003). During the latter half of the 1960s, aerospace contractors and

others in the defense industry applied systems techniques to social problems, including traffic

congestion, urban blight, crime, poverty, housing shortages, and environmental pollution.98

Though expectations were high, the actual results were mixed.99 Edwards charts a similar

migration of the systems approach in the work of Jay Forrester, an electrical engineer who

developed the Whirlwind computer, the first real-time digital computer and a crucial part of the

SAGE (Semi-Automatic Ground Environment) real-time defence system intended to detect and intercept Soviet aircraft entering US airspace. After this military work, Forrester accepted an appointment to MIT’s Sloan School of Management, where he started modeling different complex entities using a similar modeling framework. Forrester called this framework . He began applying it to industrial production, and later expanded to model urban settings, and eventually the entire world. His system dynamics framework provided the modeling basis for the influential, if controversial, publication Limits to Growth in 1972, which

96 Hughes, Rescuing Prometheus, ch. 4. 97 Ibid. 98 Jennifer Light, From Warfare to Welfare: Defense Intellectuals and Urban Problems in Cold War America (Baltimore, MD: Johns Hopkins Press, 2003). 99 Defence intellectuals and contractors generally found these social problems much more complex than the military and technological problems they had previously focused on. See: Hughes, Rescuing Prometheus, 166-76.

64 predicted ecological collapse if checks were not placed on economic and .100

Forrester’s work is an early example of the world modeling that became common after 1972, where computer models were used to forecast long-term and large-scale trends for policy purposes.101 In 1981, modeler Harold Guetzkow observed that these models had come to form,

“an integral part of the intellectual scene throughout the world.”102

The Coalescing Semantic Void

The migration of systems approaches into the Earth sciences was part of a more general trend whereby these approaches were applied to ever more areas of scientific research and to ever more social issues (with varying degrees of success). Some of the most common systems approaches relied on mathematical and computer modeling. In the Earth sciences, this modeling began in ecology and atmospheric science and later expanded into other fields like oceanography and .103 Edwards argues that computer modeling—a subset of a systems approach— provided a necessary condition for the possibility of viewing the world as a system. These models “made sense of the [global] data; they made a coherent world from collections of bits. In

100 Edwards, “The World in a Machine,” 229-31, 236-9. One of the most controversial aspects of Forrester’s models—apart from their tendency to frequently (for critics, all too frequently) predict “overshoot and collapse” scenarios—was that they relied heavily on relationships between different feedback loops (systemic relationships) rather than on observational data. However, this may have been a practical necessity, as Edwards and others like Fernando Elichirigoity argue, since at this time there were no global observational datasets available for the variables used by the Limits to Growth models (natural resources, population, pollution, capital, and agriculture). In fact, Edwards argues that it was global models like these that drove the collection of global data. See: Edwards, “The World in a Machine,” 243-4; Fernando Elichirigoity, Planet Management: Limits To Growth, Computer Simulation, and the Emergence of Global Spaces (Evanston, IL : Northwestern University Press, 1999). 101 See: Richard Ashley, “The Eye of Power: The Politics of World Modeling,” International Organization 37, no. 3 (Summer 1983): 496-7. 102 Quoted in: Ashley, “The Eye of Power,” 496. 103 Spencer Weart, in his history of the “discovery” of global warming, notes two major obstacles oceanographers faced in their efforts to model ocean processes. First, they lacked the amount of data available for the atmosphere. Second, while atmospheric modelers could use simplified equations or averaged numbers to represent many of the complex but short lived movements of air (for instance, turbulent eddies), the decades-long movement of waters in the oceans had to be modeled and computed “in full detail.” See: Spencer R. Weart, The Discovery of Global Warming (Cambridge, MA: Harvard University Press, 2003), 135.

65

a certain important epistemological sense, they gave us ‘the world’ as an ecological and physical

unity.”104 Computer modeling facilitated an understanding of the entire Earth as a contained

entity with interlinking parts, a system of interconnected components, with processes that could

be represented mathematically, integrated into a computer model, and calculated in a reasonable

span of time. The global “contents” of these models were largely provided by Earth observing

satellites.

All of this is not to say that Earth observing satellites could actually collect data on all

parts of the Earth, whatever that would mean. There are always tradeoffs to be made between a

variety of criteria, including surface coverage extent, spatial and temporal resolution, times of

day for measurements, and ideal orbits for specific instruments, not to mention financial and risk

considerations.105 Similarly, computer modeling has its own tradeoffs and limitations that make

them always less than actually global in character.106 Whatever their practical global limitations,

satellites are certainly capable of collecting more comprehensive and long-term datasets than,

say, Humboldt. On his sole trip up Mount Chimborazo, Humboldt and his team paused at

various locations where they removed state-of-the-art precision instruments out of wooden boxes transported by mules, like a cyanometer to measure the blue of the sky and a Saussure

104 Edwards, “The World in a Machine,” 242. 105 Beyond tradeoffs, there are other issues with satellite data that also require consideration, including distortions from atmospheric optical effects and orbital drift/decay, loss of instrument calibration over time, and the need to validate satellite measurements with ground-based measurements (called “ground truthing”). The ESSC’s second report describes many of these satellite issues. See: Earth System Sciences Committee (ESSC), Earth System Science: A Program For Global Change: A Closer View (Washington, DC: NASA, 1988), ch. 7. 106 ESSC, Closer View, 95. Edwards describes one of the most problematic aspects of modeling Earth processes as what modelers call parameterization. Parameterization is carried out when physical processes are too small or too complex or too little understood to be resolved in models by the laws of physics. However, these processes still need to be included as mathematical representations to obtain as realistic results as possible. This is often completed in a trial by error manner, with modelers tweaking the parameterizations as necessary. See: Edwards, A Vast Machine, ch. 13.

66

hygrometer to measure the amount of water vapour in the atmosphere.107 All aspirations aside,

Humboldt’s pointillistic observations could never match the more synoptic, continuous coverage

offered by satellites.108 There was also no way to analyze collected data via numerical

calculations other than laboriously by hand. Richardson’s early attempt at NWP illustrates the

limits of manual data analysis.

Scientists considered satellite data (and GCMs) global not because this data was collected

for every conceivable aspect of the planet (or that a GCM could represent the full complexity of

the planet), but because satellites and computer models can have a global reach in a Latourian

sense. Bruno Latour compared the global reach of scientific knowledge to a railroad. The

railroad, like all technological networks, begins locally. It consists of individual, particular

points, but these points can be expanded to reach the entire globe. Though it may not actually

touch every single spot on Earth, and certain instruments or apparatus are still required in order

to gain access, the scope of a technological network can be global. In the same way that a

railroad network can be considered global in scope despite not literally touching all spots on the

planet or being equally accessible to all, so too can satellite data be global even though

observations are not collected on every single planetary component.109 Satellites can claim

“global” coverage because they traverse the globe in a way that no other observation system ever

107 Michael Dettelbach describes the controversial nature of requiring precision measurements in the late eighteenth and early nineteenth century, as practitioners fought over whether more precise instruments measured variations in natural phenomena or simply recorded unnecessary noise and complicated later analyses. See: Michael Dettelbach, “The Face of Nature: Precise Measurement, Mapping, and Sensibility in the Work of Alexander von Humboldt,” Studies in History and Philosophy of Science Part C 30, no. 4 (1999): 473-504. Humboldt’s other measuring instruments and equipment included: barometers, chronometers, thermometers, sextents, microscope, surveying chains, balance, rain-gauge, magnetometers, dipping needles, evaporating pan, plant presses, dissecting tools, chemical and galvanic equipment, and various small repair tools. See: Stephen T. Jackson, “Instruments Utilized in Developing the Tableau physique,” in Essay on the Geography of Plants, ed. Stephen T. Jackson, trans. Sylvie Romanowski (Chicago: University of Chicago Press, 2008), 221-6. 108 Dorothy Harper, Eye in the Sky: Introduction to Remote Sensing, second edition (Montréal: Multiscience Publications, 1983), 4, 42, 140. 109 Bruno Latour, We Have Never Been Modern, trans. Catherine Porter (Cambridge, MA: Harvard University Press, 1993), 117.

67

has. The upshot is that, as early as the 1960s, some scientists began to think of satellites, paired

with computer models, as tools for studying the planet globally. These tools were thought of as

sufficiently global to provide an empirical basis for world-spanning theories about Earth processes. With these tools, the whole planet could now be studied empirically. It could now become an object of scientific inquiry.

Humboldt and Vernadsky—through “universal science” and “the biosphere” respectively—both produced holistic theories of the Earth. They believed that their global theories required the amassing and ordering of quantitative measurements from all parts of the globe, though they also recognized that their current limited capacities to assemble this empirical

data could not adequately support their global theoretical ambitions. Their speculative ideas

gestured towards a growing but still nascent semantic void, with no contemporary word or

phrase deemed adequate for describing the whole Earth scientifically. This void increased dramatically after the development of satellites and computer modeling in the late 1950s.

Scientists believed they finally possessed the tools to study the Earth more comprehensively as a

scientific object. However, there was no single, common way to refer to such a planet. Edwards

describes this period—a period that included satellite data and computer models, but also images

of the planet from space, the world-spanning United Nations, the interdisciplinary IGY, and what

he calls Cold War “closed world discourse” that divided the finite world between two major

superpowers—as an “overdetermined semiotic web.” The combination of these various strands

supported the idea that the world was a global, interconnected “system.”110 What Edwards calls

an “overdetermined semiotic web” this dissertation calls a “semantic void,” which arguably

110 Edwards, A Vast Machine, 2-3 [emphasis added].

68

better captures the vacuum-like conditions that developed from the 1960s into the 1980s, into

which the ESSC’s Earth system concept readily flowed.

THE EARTH SCIENCES AND THE SEMANTIC VOID

In addition to the technical capacities provided by satellites and computer models, scientists also

attended conferences where they were exposed to nascent but tantalizing links between Earth

science disciplines that had previously functioned in relative isolation. NASA’s Dixon Butler—

who wrote his PhD dissertation on the ionosphere of Venus and played a seminal role in the

development of NASA’s Earth Observing System (EOS) satellites in the 1980s (chapters three

and four)—recalls attending such a session at an American Geophysical Union meeting in the

late 1970s. Butler described a presentation by soil scientist Constant Delwiche, who spoke about

microbial nitrification and denitrification in soil that provided a significant source of nitrous

oxides in the stratosphere. According to Butler, his mind was “blown” at hearing Delwiche

describe the connections between the soil and the upper atmosphere. It transformed Butler’s

thinking about the planet and the kinds of research that could (and should) be conducted.111

ESSC member Berrien Moore similarly credited a conference with providing him the

opportunity to transition from his narrow mathematical focus into broader modeling of

biogeochemical cycles. In a chance encounter, Moore met oceanographer and future NASA

Administrator Robert Frosch at a Scripps Institute oceanographic conference in 1976. Frosch

invited Moore to the Woods Hole Oceanographic Institute (WHOI), where Moore attended

lectures by Bert Bolin, a pioneer in studying the carbon cycle and the buildup of carbon dioxide

111 Dixon M. Butler, interviewed by Rebecca Wright, Washington, DC, 25 Jun 2009, NASA Johnson Space Center “Earth System Science at 20 Oral History Project,” accessed 10 Oct 2019, https://historycollection.jsc.nasa.gov/JSCHistoryPortal/history/oral_histories/NASA_HQ/ESS/ButlerDM/ButlerDM _6-25-09.pdf.

69

in the atmosphere and oceans. After Frosch became NASA’s Administrator under the Carter

Administration in 1977, Frosch invited Moore to, “come down and spend some time at NASA

every once in a while. After all, they do Earth science.”112

These experiences were not limited to conferences. Institutions and international science

projects formed during this period provided venues where scientists could share these emerging

ideas about how to use global datasets to study planet-scale interconnections. The National

Center for Atmospheric Research (NCAR) and the Global Atmospheric Research Program

(GARP) provide institutional examples of these trends. In addition, some scientists working at

these institutions began developing planet-encompassing theories to explain these interconnections en masse. James Lovelock and his Gaia hypothesis was one notable example of the kind of panoramic speculation prompted by the technical changes in the Earth sciences in the mid-twentieth century.

National Center for Atmospheric Research (NCAR)

NCAR represented the culmination of research trends and the projection of future trends in the nascent field of atmospheric science. Beginning in the late 1950s, meteorology transitioned from a specialty that developed weather forecasts based primarily on observations collected from dispersed weather stations and human intuitions into a broader field of research that included any discipline interested in “,” on Earth or otherwise. During this period, atmospheric science became what James Fleming calls an “umbrella term” to denote this common research

112 Berrien Moore, interviewed by Rebecca Wright, Norman, OK, 4 Apr 2011, NASA Johnson Space Center, “Earth System Science at 20 Oral History Project,” accessed 10 Jun 2019, https://historycollection.jsc.nasa.gov/JSCHistoryPortal/history/oral_histories/NASA_HQ/ESS/MooreB/MooreB_4- 4-11.pdf.

70

interest.113 Atmospheric science emerged, according to Fleming, in the late 1950s as a

confluence of events and people: researchers searching for new levels of funding and better

access to new technologies like satellite data and computers, the maturation of theoretical models

in meteorology, and increasingly global observational networks. It also received a boost from

Cold War “angst” during the late 1950s and early 1960s generated by the uneasy feeling that the

US was falling behind the USSR in terms of weather research, which could have strategic

implications in terms of weather modification. US officials maintained that the US could not

afford to fall behind in a “weather race.”114 Practitioners in atmospheric science usually

specialized in one of many areas (atmospheric chemistry or physics, numerical weather

prediction, computer modeling, satellite meteorology, climatology [later climate science],

paleoclimatology, or upper atmospheric research) and the field had close affiliations with related

Earth science disciplines like oceanography, , bioclimatology, planetary science, and

space physics. Fleming characterized atmospheric scientists as, “a large tribe interested in

atmospheres.”115

Atmospheric science was, from the beginning, a diverse field that fostered

interdisciplinary research and collaboration via the use of increasingly global datasets and

computer models. Lloyd Berkner—a physicist, engineer, early Cold Warrior, and a major force

behind the IGY—spoke on the “Horizons of Meteorology” at a meeting of the American

Meteorological Society (AMS) in May 1957.116 Even before the launch of the world’s first

113 Fleming, Inventing Atmospheric Science, 194. 114 Fleming, Inventing Atmospheric Science, 201. For more details on the history of weather modification, see: Kristine C. Harper, Make It Rain: State Control of the Atmosphere in Twentieth-Century America (Chicago: University of Chicago Press, 2017); James Rodger Fleming, Fixing the Sky: The Checkered History of Weather and Climate Control (New York: Columbia University Press, 2010). 115 Fleming, Inventing Atmospheric Science, 194. 116 See: Allan A. Needell, Science, Cold War and the American State: Lloyd V. Berkner and the Balance of Professional Ideals (New York: Routledge, 2000).

71

artificial satellite in October of that year, Berkner was predicting that new technologies like

satellites and digital computers would drastically alter the field. According to Berkner, Earth

satellites were, “a powerful new tool that may do more to revolutionize meteorology than

anything that has happened in the last century.”117 Perhaps recognizing an opportunity to

capitalize on and facilitate the use of these new technologies, the National Academy of Sciences’

(NAS) Committee on Meteorology began advocating in November 1957 for a “National Institute

for Atmospheric Research.” By the end of 1958, NAS’s Committee on Meteorology had

changed its name to the Committee on Atmospheric Science, indicating the interdisciplinary shift

in atmospheric studies. This new committee convened the University Corporation for

Atmospheric Research (UCAR) as a non-profit consortium comprised of members from 14

universities with meteorology programs. UCAR incorporated in 1959 and formally announced a

plan to form a National Institute for Atmospheric Research to be run by UCAR.118 This institute

became the still extant National Center for Atmospheric Research (NCAR).119

From the earliest planning phases, NCAR was to be a hub of interdisciplinary research where scientists studied the atmosphere and its interconnections with land, water, biota, and the sun. In its preliminary plans for what would become NCAR, in February 1959 UCAR officials compiled a report outlining their interdisciplinary and global visions for the center. While primarily organized around the atmosphere at all altitudes (its motions, its chemical composition, its energy balance), this research would be incomplete if it focused on the atmosphere in isolation. To understand the atmosphere, according to the report, it was also crucial to study how it interacted with other components of the planet and near-space environment. NCAR would

117 Quoted in: Fleming, Inventing Atmospheric Science, 198. 118 The 14 original UCAR university members were: Arizona, Chicago, Cornell, Florida State, Johns Hopkins, Michigan, MIT, NYU, Penn State, St. Louis, Texas A&M, UCLA, Washington, and Wisconsin. See: Fleming, Inventing Atmospheric Science, 202. 119 Fleming, Inventing Atmospheric Science, 201-2.

72

include traditional laboratory facilities, but it would also run a fleet of airplanes, as well as

remote sensing and computer facilities to support more global and interdisciplinary work that

were not available at individual universities.120 Neither UCAR’s first choice for NCAR’s

director (James van Allen) or their ultimate choice (Walter Orr Roberts) were meteorology

experts. Van Allen’s research interests focused on terrestrial magnetism and space physics,

while Roberts was an astronomer and head of the High Altitude Observatory. Nevertheless,

Roberts was esteemed as a scientist and administrator with broad interdisciplinary interests and a

knack for securing government funding for research.121 These traits were clearly more valued

than meteorological specialization.

Even the design of NCAR’s home, the Mesa Laboratory, reflected a vision for the center as a focal point for interdisciplinary Earth and space science research. Roberts’ first major task after being named Director in 1960 was to hire an architect to design its headquarters on a mesa in the foothills of the Rocky Mountains in Boulder, Colorado. Echoing other scholarship that focuses on the link between architecture and science—notably Peter Galison and Emily

Thompson’s The Architecture of Science (1999)122—Stuart Leslie argues that the design of

scientific buildings can give, “concrete expression (literally) to distinctive philosophies of

research.”123 The specific design of a laboratory can, according to Leslie, provide insights into

the visions of those responsible for the design. Leslie suggests that the design of Mesa

Laboratory emerged from the iterative negotiations between Roberts and architect I.M. Pei.

Roberts’ aim was to encourage interdisciplinary collaborations and discussions via the design of

120 University Corporation for Atmospheric Research (UCAR), “Preliminary Plans for a National Institute for Atmospheric Research,” accessed 27 Oct 2019, https://opensky.ucar.edu/islandora/object/archives:3054. 121 Fleming, Inventing Atmospheric Science, 205. 122 Peter Galison and Emily Thompson, eds., The Architecture of Science (Cambridge, MA: MIT Press, 1999). 123 Stuart W. Leslie, “‘A Different Kind of Beauty’: Scientific and Architectural Style in I. M. Pei's Mesa Laboratory and Louis Kahn's Salk Institute,” Historical Studies in the Natural Sciences 38, no. 2 (Spring 2008): 174.

73

the building itself, rather than via administrative fiat. He envisioned NCAR as an

interdisciplinary center where global observational data would be assembled, analyzed, and

incorporated into computer models by NCAR staff and visiting researchers, with the Mesa

Laboratory serving as a “village green” for the community.124 Scientists would work in small

groups, but have ample opportunities for impromptu discussions with colleagues from other

groups. Rejecting what he felt were the endless corridors and sprawling wings of large

government laboratories like those of the National Bureau of Standards (also in Boulder),

Roberts emphasized a design that would promote creativity, communication, and complexity.

Pei’s final design incorporated these requirements into a structure with three five-story towers

rising out of a common two-story central structure with communal spaces (e.g. the library,

cafeteria, and meeting rooms). To support Roberts’ desire for “complexity and surprise,” Pei

kept corridors short, offices small, and grouped them around central administrative areas. With

its intertwining short corridors, there were multiple ways to get from point A to point B, which

provided Roberts with his desired, “endless opportunities for serendipitous encounters and

exchanges.”125

Not everyone agreed that Pei’s final design actually reinforced Roberts’ interdisciplinary

goals.126 Regardless, there is no doubt that interdisciplinarity was a top priority for Roberts’

NCAR, and it was a priority that continued under subsequent NCAR directors. By the time that

future ESSC chair Francis Bretherton became Director of NCAR in 1974 (chapter three), the

institution was home to a large group of permanent and rotating researchers. In particular, it had

made a name for itself in the development of GCMs and supercomputing. NCAR also hosted

124 Ibid, 178. 125 Ibid, 178-9. 126 Leslie lists a few critics, including meteorologist Robert Fleagle and chaos theorist Edward Lorenz. See: Leslie, “‘A Different Kind of Beauty,’” 177-8, 182.

74

conferences and talks on a diverse array of interdisciplinary topics, including one by James

Lovelock, who spoke about his Gaia hypothesis (see below). To support early-career scientists,

NCAR offered a number of postdoctoral research fellow positions. Eric J. Barron—an Earth

scientist who would later serve as the President of Pennsylvania State University, the President

of Florida State University, and the Director of NCAR—received undergraduate and graduate

training in geology and oceanography. By the early 1980s, Barron had developed an interest in

the relationship between plate tectonic movements and atmospheric circulation. A meteorology

professor encouraged him to apply for a postdoctoral position at NCAR, which Barron initially

found “hysterical...because there was no way a geologist was going to go to the National Center

for Atmospheric Research.” However, much to Barron’s surprise, NCAR offered him a position.

Barron later learned that NCAR awarded at least one “oddball” fellowship each year, to someone

working at the intersection of atmospheric science and other Earth science disciplines.127 With

opportunities like those afforded by postdoctoral positions, NCAR promoted research that

worked at the interstices of traditional Earth science disciplines. By the mid-1980s, it had fully

embraced this position. In 1986, UCAR created the non-profit Office for Interdisciplinary Earth

Studies (OIES) that supported research traversing the regular Earth science disciplines by

organizing workshops, conferences, and study sessions, as well as publishing reports, brochures,

and newsletters.128

127 Eric J. Barron, interviewed by Rebecca Wright, Tallahassee, FL, 1 Jul 2010, NASA Johnson Space Center “Earth System Science at 20 Oral History Project,” accessed 27 Oct 2019, https://historycollection.jsc.nasa.gov/JSCHistoryPortal/history/oral_histories/NASA_HQ/ESS/BarronE/BarronE_7- 1-10.pdf. 128 Prime Contract No. NAS5-29372, Schedule A, [1987], Folder 114, Earth System Sciences Collection, NCAR, Boulder, CO; Letter from Payson R. Stevens to Laura Lee McCauley, 15 Jul 1987, Folder 156, Earth System Sciences Collection, NCAR, Boulder, CO; “Office for Interdisciplinary Earth Studies (OIES) Records,” NCAR Archives, accessed 28 Oct 2019, https://opensky.ucar.edu/islandora/object/archives%3A8777.

75

Global Atmospheric Research Program

The Global Atmospheric Research Programme (GARP) formed at the confluence of satellites

and computer modeling and facilitated empirical studies of the planet in more holistic and

interconnected ways. The United Nations, with impetus from a report to US President John

Kennedy about the decadal prospects for atmospheric science, passed two resolutions that called

for international cooperation in the development of improved weather forecasts, research on the

weather and climate, and a global observation system.129 In response to the research component

of the UN call and building off the success of the IGY (1957-8), the International Council of

Scientific Unions (ICSU) and the World Meteorological Organization (WMO) jointly created

GARP to organize and lead an integrated research program in 1967. Broadly, GARP would

conduct research projects on the physical processes in the lower and upper atmosphere that

contributed to weather and climate using Earth observing satellite observational capacities along

with other in-situ methods, as well as computer modeling. These tools would provide evidence

to show that meteorological events could only be understood if the atmosphere was studied, “as a

single physical system where every part is interacting with every other part.”130 By

understanding the atmosphere as a whole, researchers believed that they could better understand

and forecast local conditions. In 1969, GARP’s structure and tasks were finalized. It would

conduct a number of smaller field experiments including one in the tropics called GATE (GARP

129 These are UN General Assembly Resolution 1721 (XVI) adopted on 20 December 1961 and Resolution 1802 adopted on 14 December 1962. 130 Global Atmospheric Research Programme (GARP) Joint Organising Committee (JOC), An Introduction to GARP (GARP Publication Series No. 1, October 1969), 15.

76

Atlantic Tropical Experiment), a Data Systems Test (DST) to evaluate the global observing

system, and the First GARP Global Experiment (FGGE).131

GARP’s primary goal was FGGE—also known as the Global Weather Experiment—that would produce global datasets to improve numerical models of the weather and climate. At the end of 1978 and continuing into 1979, FGGE collected data from five geostationary and four polar orbiting satellites, as well as hundreds of ships, drifting buoys, aircraft, and constant-level

balloons. Edwards describes GARP and FGGE as a “quantum jump for meteorology” in terms

of the sheer scale of data collected as well as its unprecedented coverage of the southern

oceans.132 While the IGY was arguably a more comprehensive program with respect to the kinds

of data it collected and the number of Earth science disciplines it incorporated (basically all of

them apart from the environmental sciences), GARP’s FGGE compiled vastly more data and did

so in a usable format. According to Elena Aronova, the IGY data regime emphasized the

accumulation of as much data as possible rather than facilitating immediate use. The data were,

therefore, stored using the most durable storage technology of the time: miniaturized analog

images shrunken onto microfilm.133 In contrast, FGGE data were collected and stored in digital format on magnetic tape, which was a more fragile storage medium but made the data immediately accessible by digital computers. This data represented the first global, usable dataset for meteorology.134 In 1980, GARP transitioned into the World Climate Research

Program (WCRP) to focus more specifically on climate-related issues that went beyond the

atmosphere to incorporate the oceans and land surfaces. It conducted a number of

interdisciplinary projects in the 1980s that included the Ocean Drilling Program, the World

131 Conway, Atmospheric Science at NASA, 70. 132 Edwards, A Vast Machine, 249. 133 Aronova, “Geophysical Datascapes of the Cold War,” 307-27. 134 Conway, Atmospheric Science at NASA, 65; Edwards, A Vast Machine, 249.

77

Ocean Circulation Experiment (WOCE), and the International Satellite Land Surface

Climatology Project (ISLSCP).135

The GARP Atlantic Tropical Experiment (GATE) exemplified the more global and

interdisciplinary Earth science research developed and undertaken in the 1960s and 1970s,

driven in part by the newly available technologies. GATE was undertaken to better understand

how solar heat stored in tropical oceans contributed to global air circulation. Better

understanding meant improved numerical models of these processes. From June to September

1974, 20 countries participated in this field experiment that incorporated nearly 1,000 land

stations, 38 research ships, 65 moored buoys, 13 aircraft, and six satellites along a wide section

of the Atlantic Ocean and coastal areas, from the east coast of South America to the west coast of

Africa. From these platforms, scientists collected a range of data on the ocean-atmosphere

boundary layer: air and sea surface temperature, pressure, humidity, wind speeds, precipitation,

solar and infrared radiation, water salinity, current speed, and dust layers in the air, along with

the liquid water content, height, temperature, motion, and speed of clouds.136 According to

Edwards, the amount of data collected during GATE vastly exceeded the amount normally

available for weather forecasting and so, given available computational power, processing took

almost three years. Finished datasets were not available until 1977.137 Though GATE focused

solely on the physical properties of the ocean-atmospheric boundary layer, historian of science

Spencer Weart suggests that this style of research project encouraged further interdisciplinary

135 Conway, Atmospheric Science at NASA, 92, 233-4; ESSC, Closer View, 154-5. 136 R.J. Polavarapu, and G.L. Austin, “A Review of the GARP Atlantic Tropical Experiment (GATE),” Atmosphere- Ocean 17, no. 1 (1979): 3-6. 137 Edwards, A Vast Machine, 245.

78

forays, as scientists looked to study other planetary elements—for instance biotic elements—that

could influence atmospheric conditions.138

A 1975 publication by the US Committee for GARP captured prevailing scientific

sentiments about the possibility of a scientifically robust conception of the planet as an

interconnected system. In Understanding Climatic Change: A Program for Action (1975), the

US Committee for GARP—whose membership included future ESSC chair Francis Bretherton—

recommended a US climatic research program to improve knowledge of the basic forces

affecting the “global climate.” The committee stressed that, “there is a new generation of

atmospheric scientists. Their tools are the computer, numerical models, and satellites, and they

know how to use them well.”139 Interestingly, the committee also argued that, to understand the

“global climate,” it was not enough to simply study the atmosphere and the oceans. Other parts

of the planet also required attention. The committee presented its understanding of the climate in

an abstract diagram of a global “coupled atmosphere-ocean-ice-earth climatic system” (Figure

1.1).140 For the committee, the climate was now best thought of as a system comprised of the

atmosphere, hydrosphere, cryosphere, lithosphere, and biosphere, where physical, chemical, and

biological processes interacted to create the planet’s climatic conditions.141 Unlike the ESSC’s

Earth system “wiring” diagram (chapter three), the US Committee on GARP’s representation

contained pictographic depictions of some physical features of the Earth like waves, clouds,

mountains, and ice caps. Arrows represented more abstract flows of material and energy in the

138 Weart, The Discovery of Global Warming, 100. 139 United States Committee for the Global Atmospheric Research Program (GARP), National Research Council, Understanding Climatic Change: A Program for Action (Washington, DC: National Academy of Sciences, 1975), v- vi. 140 A near-dentical diagram appeared in the GARP Joint Organising Committee’s 1975 publication on The Physical Basis of Climate and Climate Modelling. See: Global Atmospheric Research Programme (GARP) Joint Organising Committee, The Physical Basis of Climate and Climate Modelling, Report of the International Study Conference in Stockholm, 29 July-10 August 1974 (GARP Publication Series No. 16, April 1975), 13-14. 141 US Committee for GARP, Understanding Climatic Change, 13-16.

79

system. It was a global and interdisciplinary view of Earth and, arguably, a prototype of the

ESSC’s wiring diagram. It represented the planet as a series of interconnecting components

which could not meaningfully be studied in isolation.

Figure 1.1. The coupled atmosphere-ocean-ice-earth climatic system. (US Committee for GARP, Understanding Climatic Change, 1975)

Gaia

Beyond institutions and international organizations like NCAR and GARP, individual scientists

also proposed theories that gestured at a new scientific conception of the whole Earth. James

Lovelock’s discipline-transcending Gaia hypothesis is the most notable example during the

1970s and early 1980s. Although it captured the popular imagination, resistance to certain

interpretations kept the Gaia hypothesis from gaining widespread traction in the scientific

community. A simple definition of the Gaia hypothesis—that the Earth is a self-regulating

system comprised of feedback loops that contribute to planetary conditions—might be

uncontroversial to many, if not most, Earth and environmental scientists. But there were also

80

more controversial interpretations, the most contentious of which described biota as purposeful

controllers, keeping the Earth in a state optimal for life. The Gaia hypothesis did not gain

widespread popularity and support in scientific communities because of these controversial

teleological interpretations that could not be easily expunged from the hypothesis (chapter five).

Despite this, the hypothesis represents an early attempt by an individual to formulate a holistic

theory about the entire planet at a time when new tools and techniques offered the promise of

global data and analysis to empirically support these claims. The Gaia hypothesis received

widespread scientific attention by the late 1970s and 1980s, even if not widespread acceptance.

When Lovelock first publicly proposed his Gaia hypothesis in the 1970s, he was already

an established scientist and inventor, though with an irregular career path.142 After early training

in chemistry, Lovelock received a PhD in medicine from the University of London in 1949. He

worked for the UK’s National Institute for Medical Research for a number of years, but

eventually became an independent scientist, taking on contracts with various organizations,

including NASA and Shell Limited. Lovelock made an early name for himself as an inventor,

designing a variety of detectors for separating and measuring components of gases. The most

well-known for its usefulness was the electron capture detector (ECD) that can measure trace

142 Historical sources on James Lovelock and the Gaia hypothesis are sparse. There is a biography by John and Mary Gribbin that provides useful details about Lovelock’s life, though it presents a rather whiggish view of the Gaia “hypothesis”—the Gribbins claim that it transitioned from an unsubstantiated hypothesis to an empirically proven theory—as being the truth “out there” about the planet that Lovelock “discovered” without any recognition of the constructed nature of science. Michael Ruse’s The Gaia Hypothesis: Science on a Pagan Planet (2013) represents the most scholarly treatment, albeit more philosophical than historical. Lovelock himself has also provided historical details about Gaia in his many books on the subject. See: John Gribbin and Mary Gribbin, James Lovelock: In Search of Gaia (Princeton, NJ: Princeton University Press, 2009); Michael Ruse, The Gaia Hypothesis: Science on a Pagan Planet (Chicago: University of Chicago Press, 2013); J.E. Lovelock, Gaia: A New Look at Life on Earth (New York: Oxford University Press, 1987 [1979]); James Lovelock, Homage to Gaia: The Life of an Independent Scientist (New York: Oxford University Press, 2000); James Lovelock, The Revenge of Gaia: Why the Earth is Fighting Back - and How We Can Still Save Humanity (Toronto: Penguin Books, 2007); James Lovelock, The Vanishing Face of Gaia: A Final Warning (Toronto: Penguin Books, 2010).

81 amounts of gaseous compounds, including some atmospheric pollutants.143 It was used by researchers in Britain and the US to detect concentrations of pesticides like dichlorodiphenyltrichloroethane (DDT) in air samples and, according to Lovelock and his biographers, it provided the empirical evidence that led the US to almost completely ban DDT in

1972.144

In the early 1970s, Lovelock used an ECD to detect levels of chlorofluorocarbons (CFCs) away from industrial locations like on mainland Britain where they would have been expected.

Unexpectedly, CFCs were detected throughout the remote reaches of the Southern Atlantic

Ocean and even in Antarctica. These measurements demonstrated the pervasive nature of human pollution across the planet, and they in part led to research on the role of CFCs in stratospheric ozone depletion undertaken by Mario Molina and Sherry Rowland.145 In 1974, Lovelock was elected to the Royal Society of Fellows. His Fellow citation demonstrates his eclectic background that also included the study of the transmission of respiratory infection, methods of air sterilization, and cryogenics. It described Lovelock’s work as showing, “remarkable originality, simplicity and ingenuity.”146

143 The ECD works by using beta rays to remove electrons from nitrogen gas. These free electrons are attracted to a positively charged electrode, which produces an electrical current. Molecules from introduced gases absorb these free electrons and reduce the electrical current. Since every chemical compound has a different “affinity” for electrons, these differences allow for the identification of different compounds based on absorption rates. The ECD was so sensitive that it could, according to the Gribbons, “detect the presence of a couple of hundred thousand molecules of DDT in a cubic centimeter of air. This amounts to one-tenth of a femtogram of DDT, and one femtogram is one-billionth of a millionth of a gram.” See: Gribbon and Gribbon, James Lovelock, 97; James E. Lovelock, “A Sensitive Detector for Gas Chromatography,” Journal of Chromatography 1 (1958): 35-46. 144 Lovelock claims that, while Rachel Carson’s work receives the majority of attention, it was actually his device that led to the widespread ban. See: Gribbon and Gribbon, James Lovelock, 118-9. 145 Gribbon and Gribbon, James Lovelock, 121-31; Mario J. Molina, and F.S. Rowland, “Stratospheric Sink For Chlorofluoromethanes: Chlorine Atom Catalysed Destruction of Ozone,” Nature 249, no. 5460 (28 Jun 1974): 810- 2. 146 “Citation: James Ephraim Lovelock, 1974” The Royal Society, accessed 4 Nov 2019, https://archive.ph/20140410000832/http://royalsociety.org/DServe/dserve.exe?dsqIni=Dserve.ini&dsqApp=Archive &dsqDb=Catalog&dsqCmd=show.tcl&dsqSearch=(RefNo==%27EC/1974/16%27).

82

While respected for certain scientific contributions, Lovelock’s Gaia hypothesis proved

highly controversial. Lovelock claimed that one source of the hypothesis could be traced back to

his time as a contractor with NASA’s Jet Propulsion Laboratory (JPL) in the 1960s. This root

may link the Gaia hypothesis to the “overview” perspectives offered by satellites. At JPL,

Lovelock worked on life detection instruments and experiments intended for NASA’s Voyager

Mars Program, which transformed into the Viking Program that sent two probes to Mars in the

mid-1970s. Recognizing the difficulties of envisioning what extraterrestrial life might look like,

let alone recognizing it, Lovelock proposed using simple experiments on atmospheric

composition to detect the chemical signatures of life rather than looking for specific

extraterrestrial lifeforms. Lovelock presumed that the atmospheres of “dead” planets would look

much different than those teeming with life. According to Lovelock, the atmosphere of a planet

with no life would, after a certain period of time, reach a state of chemical equilibrium. These

are the observable conditions on planets like Mars and Venus, with atmospheres largely

composed of non-reactive gases like carbon dioxide. If, however, one observes an atmosphere in

chemical disequilibrium, containing significant amounts of reactive gases like oxygen and

methane, one should assume the presence of a continual source for these gases. An obvious

continual source are the metabolic processes of biota as observed on Earth. Simply by analyzing

a planet’s atmospheric composition, one could identify the signatures of life or its absence.147

Building on these ideas, Lovelock developed the Gaia hypothesis gradually throughout the 1960s. By 1970, he was collaborating with microbiologist Lynn Margulis.148 Margulis has

147 J.E. Lovelock, “A Physical Basis For Life Detection Experiments,” Nature 207 (7 Aug 1965): 568-70; Dian R. Hitchcock, and James E. Lovelock, “Life Detection By Atmospheric Analysis,” Icarus 7 (1967): 149-59. 148 Margulis claimed that “at least four different scientists” suggested that she contact Lovelock. In his research on the early relationship between Margulis and Lovelock, Bruce Clark, citing Lovelock, notes that Margulis had initiated correspondence with Lovelock in the summer of 1970 “due to ’s recommendation[.]” See: Bruce Clark, “Gaia is not an Organism: Scenes from the Early Collaboration Between Lynn Margulis and James

83 been described as a “scientific rebel,” a supremely confident scientist with the tenacity to stick by her ideas even in the face of great adversity from the broader .149

Arguably Margulis’ most significant scientific contribution was her promotion of Serial

Endosymbiosis Theory (SET), which suggests that evolutionary change comes not solely from gene mutation, but can also occur via symbiotic relationships. Initially controversial, SET— which explains how non-nucleated bacteria (prokaryotes) evolved into nucleated eukaryotes via the symbiotic combination of previously separate microorganisms—gained greater traction throughout the 1970s. It is now the standard explanation for the evolution of eukaryotes.150 By

1983, Margulis was sufficiently mainstream to be elected to the National Academy of

Sciences.151

Lovelock and Margulis published three papers in 1974 that introduced scientists to the

Gaia hypothesis. Their first and most substantial article, “Biological Modulation of the Earth’s

Atmosphere,” was published by Margulis’ ex-husband Carl Sagan in his role as editor-in-chief of the interdisciplinary planetary science journal Icarus.152 Lovelock and Margulis drew on Sagan

Lovelock,” in Lynn Margulis: The Life and Legacy of a Scientific Rebel, ed. Dorion Sagan (White River Junction, VT: Chelsea Green Publishing, 2012), 36. 149 The phrase “scientific rebel” comes from the title of a volume edited by Margulis and Carl Sagan’s son Dorion Sagan that appeared after Margulis’ death in 2011 to remember her scientific legacy. Though Dorion Sagan was Margulis’ son, he was also her colleague, as they had also co-authored a number of books. See: Dorion Sagan, ed., Lynn Margulis: The Life and Legacy of a Scientific Rebel (White River Junction, VT: Chelsea Green Publishing, 2012). 150 Margulis notes that acceptance of cellular mitochondria and plastids organelles as ancestors of previously separate entities is widespread, though this does not extend to all the organelles that Margulis proposes, for instance cilia. See: Lynn Margulis, The Symbiotic Planet: A New Look at Evolution (London: Phoenix, 1998), 50-64. 151 Margulis was elected to the NAS in the discipline of evolutionary biology. See. “Lynn Margulis,” National Academy of Sciences, accessed 4 Nov 2019, http://www.nasonline.org/member-directory/deceased- members/52941.html. 152 Lisa Messeri notes that Icarus, founded in 1962, “established the ‘planet’ as a central object of study, trumping disciplinary divisions in favor of gathering together a diversity of views that could make sense of the astronomical category to which Earth, in the post-Copernican world, belonged.” See: Lisa Messeri, Placing Outer Space: An Earthly Ethnography of Other Worlds (Durham, NC: Duke University Press, 2016), 5.

84

and George Mullen’s faint-sun hypothesis, which suggested that the sun’s luminosity had

increased substantially over its history.153 Despite increasing levels of solar radiation, the

Earth’s conditions had remained largely stable, a surprising fact Lovelock and Margulis hoped to

explain via their Gaia hypothesis. Adopting the Greek word “Gaia” to refer to the totality of

living things on Earth, Lovelock and Margulis defined Gaia as, “a complex entity involving the

earth’s atmosphere, biosphere, oceans and soil.”154 Gaia’s combined totality resulted in a,

“feedback or cybernetic system which seeks an optimal physical and chemical environment for

the biota.”155 In other words, the biological and non-biological components of the planet worked

together as a homeostatic system that regulated conditions like temperature, acidity, and

composition in a manner akin to an organism’s ability to regulate temperature. Lovelock and

Margulis argued that living organisms drove this planetary control system, and provided

examples of how biota could act as regulators by responding to changing conditions in ways that

would maintain conditions conducive for life. For example, organisms might control the planet’s

average temperature by the absorption of the greenhouse gas carbon dioxide during

photosynthesis or its emission during respiration. While Lovelock and Margulis began by

suggesting that the Gaia hypothesis was mainly a heuristic device rather than a “true description”

of the planet, later works argued more forcefully for the hypothesis, suggesting that natural

selection favoured organisms that more effectively controlled their environments.156 Lovelock

153 Carl Sagan and George Mullen, “Earth and Mars: Evolution of Atmospheres and Surface Temperatures,” Science 177, no. 4043 (7 Jun. 1972): 52-6. 154 Lovelock’s neighbour in Bowerchalke in southern England, the novelist William Golding, suggested the word “Gaia” for this hypothesis. See: Lovelock, The Vanishing Face, 128-9. 155 Lynn Margulis, and J.E. Lovelock, “Biological Modulation of the Earth’s Atmosphere,” Icarus 21 (1974): 473. 156 James E. Lovelock and Lynn Margulis, “Atmospheric Homeostasis By and For the Biosphere: The Gaia Hypothesis,” Tellus 26 (1974): 2-10; James E. Lovelock and Lynn Margulis, “Homeostatic Tendencies of the Earth’s Atmosphere,” Origins of Life 5 (1974): 93-103.

85

eventually promoted the idea that Gaia (i.e. the whole planet) was an organism because of its

apparently self-regulating homeostatic properties.157

Though not trained in the Earth sciences nor actively engaged in research that used satellite data or global computer models, Lovelock moved among scientists engaged in or knowledgeable about these new tools.158 He encountered these scientists at JPL in the 1960s,

and later at NCAR in the 1970s. Lovelock described NCAR as, “an entrancing place of science

perched on a mountainside at Boulder in Colorado.”159 At NCAR, Lovelock discussed the Gaia

hypothesis with atmospheric scientists, climatologists, and other geophysicists, including Steven

Schneider (chapter five). Lovelock’s NCAR connections were strong enough that he was invited

to give the keynote address at NCAR’s twenty-fifth anniversary in 1985. This address was,

according to NCAR Director Wilmot Hess, “one of the high points of this anniversary

celebration.”160 At the “well attended” talk, Hess effused praise for Lovelock’s scientific and

engineering achievements.161 Hess described a conversation with the atmospheric chemist Ralph

Cicerone regarding the fact that, “the atmospheric composition of the earth is dominated by biology” and would be radically different without life, which Hess took to be trivially true, saying, “I believe that of course.” The exchange resulted in Cicerone giving Hess a copy of

Lovelock’s Gaia: A New Look at Life on Earth (1979). Hess called the Gaia hypothesis, “very

157 Margulis disagreed with Lovelock on this point. For Margulis, the planet was not an organism but the product of a number of different interacting ecosystems. Gaia was simply “the largest ecosystem on Earth.” See: Margulis, Symbiotic Planet, 152; Clark, “Gaia is not an Organism,” 32-43. 158 In the 1980s, Lovelock developed a simplified computer model called “Daisyworld” to demonstrate how the physical and chemical properties of organisms (in this case light and dark daisies) could unintentionally maintain planetary conditions in a range suitable for life. See: Andrew J. Watson and James E. Lovelock, “Biological Homeostasis of the Global Environment: The Parable of Daisyworld,” Tellus B: Chemical and Physical Meteorology 35, no. 4 (1983): 284-9. 159 Lovelock, The Revenge of Gaia, 5. 160 James Lovelock, “Lecture Transcript,” National Bureau of Standards, UCAR/NCAR Oral History Collection, NCAR Archives, accessed 12 Jun 2019, http://n2t.net/ark:/85065/d747485v. 161 “Fooling Mother Nature: Lovelock Gives 25th Anniversary Lecture,” Staff Notes: National Center for Atmospheric Research 20, no. 39 (27 Sep. 1985), 4.

86

provocative and controversial and thought-provoking.” In 1985, NCAR was in the process of planning atmospheric chemistry research for the next decade, with an emphasis on understanding how, “the system works as a whole.” Hess lamented that, “We should have read this book earlier,” implying that Hess saw NCAR’s future research as being in alignment with at least some components of the Gaia hypothesis, even if specific details might be in dispute.162

Few geophysicists or atmospheric scientists or climatologists found the Gaia hypothesis

wholly appealing, but they were open to debating the ideas.163 And Lovelock was well enough

regarded by NCAR scientists to be invited to give an anniversary lecture for the center. His

ideas were not ridiculed by the audience but were taken seriously during the subsequent question period. Those in the biological sciences, however, were more hostile to the hypothesis. Many biologists labeled the Gaia hypothesis as pseudoscience, openly rejecting it with often surprising intensity. Michael Ruse attributes these strong reactions to a combination of factors. Biology was, according to Ruse, in a state of turmoil in the late 1970s and early 1980s as new techniques and approaches generated disciplinary controversy and in-fighting. At unsure times like these,

scientists are more apt to aggressively defend their disciplines and label alternative viewpoints

“pseudoscience.”164 It did not help that Gaian ideas were quickly adopted and promoted by many “new age” religious figures and other non-scientists.165 Most damning for biologists were

Lovelock’s teleological descriptions of the Earth as “Gaia” that identified the planet as an

“organism” with a “purpose,” to maintain the planet in a state optimal for life. If this was simply

a metaphor, biologists might not have reacted so harshly. However, Lovelock frequently

162 James Lovelock, “Lecture Transcript.” 163 Lovelock, The Vanishing Face, 108. 164 Ruse, The Gaia Hypothesis, 205-15. See also: Michael D. Gordin, The Pseudoscience Wars: Immanuel Velikovsky and the Birth of the Modern Fringe (Chicago: University of Chicago Press, 2012). 165 Ronald Kline makes a similar argument for why the term “cybernetics” fell out of favour by the 1970s due to its uptake by pseudoscientific groups like Scientologists. See: Kline, The Cybernetics Moment, ch. 3.

87

stressed that the Earth was not like an organism but was an organism. Many biologists rejected

these teleological declarations. They were all too similar to disreputable ideas like vitalism that

had only recently been expunged from mainstream biology.166 Ruse argues that these factors

help explain the backlash against Gaia by biologists: “when Gaia came bumbling in, it was seen

as not just wrong but radically upsetting.”167 These negative reactions, along with the variety of

interpretations offered by the Gaia hypothesis, help explain why it never transformed into a full-

fledged or widespread research program.168 Nevertheless, “Gaia” emerged during this period as

one of the many ways to refer to an interconnected planet.

CONCLUSION

NCAR, GARP, and Lovelock’s Gaia hypothesis illustrate the kinds of interdisciplinary and

globally-oriented research and ideas facilitated in part by the new tools of satellites and computer

modeling in the 1970s and early 1980s. These new tools, and the nascent research they

supported, were so different from what was previously available that many viewed them as

different not just in degree, but in kind. Recall William Nierenberg’s claim that the amount and

kinds of data that would become available to oceanographers would be, “completely foreign to

the experience of the existing academic oceanographic community.”169 Satellites and computers offered revolutionary ways to study not just regions of the planet, but the entire globe. As

166 Evolutionary biologists, biogeographers, systematics, and social behaviourists all argued amongst themselves and with each other. These disagreements were often nasty in tone. See: Ruse, The Gaia Hypothesis, 205-15. 167 Ruse, The Gaia Hypothesis, 213. 168 While the Gaia hypothesis never became a formal research program, Lovelock and a few fellow scientists did undertake some empirical studies of certain potential mechanisms whereby biological entities provided that maintained stable (and biologically favourable) planetary conditions, most notably research conducted on the role of dimethyl sulphide in cycling sulphur from the oceans back onto the land. See: J.E. Lovelock, R. J. Maggs, R. A. Rasmussen, “Atmospheric Dimethyl Sulphide and the Natural Sulphur Cycle,” Nature 237 (23 Jun. 1972): 452-3. 169 Quoted in: Conway, “Drowning in Data,” 139-40 [emphasis added].

88 satellite data and computational power increased, so did the scope of what might be integrated into Earth science models. This widened the ambit of disciplines that might meaningfully collaborate to produce these models. Researchers could not only conceive of but actually begin to study the Earth in systemic ways, treating the various components of the planet—the air, land, water, and biota—as ultimately interconnected and in need of interdisciplinary study. The Earth could now be an object of scientific inquiry in a way it could not for Humboldt or Vernadsky.

Arguably, a semantic void developed for Earth science communities during this period.

However, the semantic void depicted in this chapter differs from the one Leo Marx described.

Writers and politicians in the late nineteenth and early twentieth century actively declared that

“machines” or “mechanical arts” were inadequate to capture the changing technical landscape.

Earth scientists from the 1960s to the 1980s did not explicitly note any linguistic deficiencies.

What marks this “void” is not a complete lack of phrases to describe an interconnected planet, but a multitude of options and a lack of community consensus. Phrases like the “coupled atmosphere-ocean-ice-earth climatic system” or “Gaia” gestured towards this new conception of an interconnected planet but failed to gain widespread traction, either because of their unwieldy character or their problematic connotations. It was only after the ESSC’s development and promotion of an Earth system science research program that Earth scientists began using a single common phrase. That there was an abundance of terminology before and a consensus after the

ESSC’s work suggests that the “Earth system” filled an inchoate and unstated, yet nevertheless tangible, void. Chapters three, four, and five show how the usefully vague “Earth system” ultimately became the favoured way for Earth and environmental scientists to refer to the entire planet as an interconnected scientific object. Before that achievement, chapter two examines an early attempt by NASA and affiliated researchers to develop a large-scale research program that

89 would incorporate these new research tools to study the physical, chemical, and biological processes of the world’s lands, oceans, and atmosphere as a single, integrated system.

Chapter 2 The Life and Death of NASA's Global Habitability Initiative

INTRODUCTION

Satellites and computer models afforded Earth scientists the instrumental capacity to study the

planet in more interconnected ways. But these were not simple and inexpensive capacities to

develop and deploy. It is, therefore, not surprising that a US institution—and not an individual—

was a significant driving force behind the first major attempt to create a global research initiative

that would study the entire Earth as a single object of scientific inquiry. In 1982, 49 US

scientists convened at a NASA-sponsored workshop to negotiate the details of such an initiative.

The proposed research program—called Global Habitability—was motivated by the recognition

that the Earth had, “entered a unique epoch when one species, the human race, has achieved the

ability to alter its environment on a global scale.”1 The initiative’s proponents had an ambitious

global vision for the Earth and environmental sciences: to study the physical, chemical, and

biological processes of the world’s lands, oceans, and atmosphere as a single, integrated system.

Using a fleet of current and future Earth observing satellites for data collection, as well as

complementary in-situ observations obtained through “international collaboration,” the proposed initiative would attempt to better understand the natural and human changes affecting the planet’s living conditions.2 NASA personnel and affiliated scientists specifically developed

Global Habitability for introduction to the international community at a major international space

1 Jet Propulsion Laboratory (JPL), Global Change: Impacts on Habitability: A Scientific Basis for Assessment ([Washington, DC]: NASA, 1982), iii. That this language resembles language used by Paul Crutzen in his articles on the Anthropocene in 2000 and 2002 is no coincidence. Crutzen was one of the 49 scientists at the Global Habitability workshop. See: Paul J. Crutzen and Eugene F. Stoermer, “The ‘Anthropocene,’” IGBP Newsletter 41 (May 2000): 17–8; Paul J. Crutzen, “Geology of Mankind,” Nature 415, No. 3 (3 Jan. 2002): 23; JPL, Global Change, 15. 2 JPL, Global Change, 13.

90 91

conference, the second United Nations Conference on the Exploration and Peaceful Uses of

Outer Space (UNISPACE ‘82).

Global Habitability represents the first major institutional attempt to develop a

comprehensive Earth science research program that leveraged the new technologies of satellites

and computer modeling. No longer focused only on the atmosphere or the planet’s climate like

the research conducted as part of GARP, Global Habitability would equally prioritize all

components of the planet. Given NASA’s institutional strength in space technologies like satellites and its early work in global modeling, it was well situated to develop and conduct such a program.3 This initiative was, in part, an attempt to garner new research funding by proposing

a research program that drew on NASA’s technical strengths but also expanded the agency’s

Earth science activities. NASA sought to build a community of interdisciplinary Earth scientists

supportive of, and reliant on, the kinds of space technologies it excelled at producing.

With Global Habitability, NASA attempted to transform the Earth, the entire Earth, into

an object of scientific inquiry. The Global Habitability initiative would transform the various

physical processes of the planet into an interconnected system that could be studied as a whole

using satellites and modeled on computers. It was the first major foray into transforming the

whole Earth into a scientific object by bringing together a number of conceptual threads,

domains of scientific inquiry, institutional needs, and international opportunities. If it had

succeeded as an initiative and gained widespread traction, it is possible that Earth scientists today

might have used the phrase “global habitat” to refer to the Earth as a scientific object, rather than

the “Earth system.”4

3 Paul Edwards, A Vast Machine: Computer Models, Climate Data, and the Politics of Global Warming (Cambridge, MA: MIT Press, 2010). 4 This is simply to suggest that there is nothing inevitable about the phrase “Earth system” rather than a strong counterfactual claim regarding the phrase “global habitat.” Arguably, “global habitat” might have provided too

92

Despite NASA’s lofty goals, Global Habitability never became an actual program and has now largely been forgotten. With its global environmental focus and the possibility for international participation, NASA and US representatives were confident that the initiative would be favourably received by other countries at UNISPACE ‘82. However, much to their surprise, the response to Global Habitability was not only unfavourable, but some US conference participants reported that the initiative actually provoked open hostility. Chapter two examines the unique historical moment in which Global Habitability was both conceived and rejected to illuminate the challenges involved in transforming something into a scientific object. It was not obvious or necessary that the whole Earth should be studied in this way. It would take work to make this specific way of studying the planet compelling to scientific and political communities outside of NASA. In this particular case, the work was unsuccessful.

Though Global Habitability was peaceful and environmentally-focused, it was also a hastily and unilaterally developed initiative that would involve massive satellite data collection but incorporated no clear data policy. The scientists and institutions in chapter one understood

“data” as simply numerical observations that could feed into or help validate computer models.

There was an unstated presumption that the collection of this data was an unquestionable “good”

(or at the very least “neutral”) since it was part of the supposedly objective pursuit of scientific knowledge. This sense of empirical data is roughly equivalent to that presented in older, positivist philosophies of science.5 However, scholarship in the history of science—most recently by those using “data” as a category of historical analysis—shows that the collection and

much specificity, and therefore less utility, for it to become the predominant way for Earth scientists to refer to the planet as a scientific object today. 5 See: Carl Hempel, Philosophy of (Upper Saddle River, NJ: Prentice Hall, 1966); Karl R. Popper, Conjectures and Refutations: The Growth of Scientific Knowledge (New York: Basic Books, 1965).

93

dissemination of data is “inherently political.”6 Since it requires making choices about what

kinds of data to collect and who may (or may not) access and therefore benefit from that data,

Elena Aronova, et al. argue that the requirements of data collection inherently create hierarchical

power dynamics.7 Though always political, W. Patrick McCray notes that some datasets (like

census or climate data) carry more national security and policy implications than others (for

instance, astronomical data).8 As such, certain kinds of data generate more political “friction”

than others.9 Satellite data collected from Earth observing platforms have, at certain historical

periods, encountered high levels of political friction. While satellites could facilitate global

research opportunities for scientists, they could also be economic and political liabilities,

particularly for those in developing countries in the early 1980s who only had mediated access to

satellite data. The same space surveillance technologies that were increasingly being used to

collect environmental data to facilitate improved weather forecasting, disaster relief, or natural

resource management for a country could also be used to disclose to foreign agencies and transnational corporations crucial information about strategic natural resources, commodity futures, and national infrastructure.10

Global Habitability was the product of a small group of privileged US scientists who

failed to grasp the inherently political nature of data and the importance of the geopolitical

context of the 1970s and early 1980s. At this historical moment, less developed countries of the

6 For a recent example of the adoption of “data” as a category of historical analysis in the history of science, see: Special Issue: Data Histories, Osiris 32, no. 1 (2017). 7 Elena Aronova, et al., “Introduction: Historicizing Big Data,” Osiris 32, no. 1 (2017): 7. 8 W. Patrick McCray, “The Biggest Data of All: Making and Sharing a Digital Universe,” Osiris 32, no. 1 (2017): 244-5. 9 Paul Edwards developed the concept of “data friction” to describe the economic, social, political, and logistical difficulties involved in the collection and utilization of standardized large-scale datasets to facilitate calculations, often for numerical modeling. See: Edwards, A Vast Machine, 80. 10 Simone Turchetti and Peder Roberts, eds., The Surveillance Imperative: geosciences during the Cold War and beyond (New York: Palgrave Macmillan, 2014).

94 world (including many former colonies) coordinated their actions on the world stage to fight for more equitable economic relations and national sovereignty protections, endorsing calls for a

New International Economic Order (NIEO) and a New World Information Order (NWIO). As a result, tensions emerged between the desire for ready access to satellite data and the desire to limit this access in order to protect sovereignty interests in the wake of decolonization. In the sparse extant historiography that mentions Global Habitability, scholars like Erik Conway and

W. Henry Lambright note that the initiative received a poor reception at UNISPACE ‘82 and that countries raised concerns regarding access to and the cost of satellite imagery, but they do not elaborate on the international political and economic contexts in which these criticisms were raised.11 Without understanding these contexts, international objections to this proposed global

Earth science research program are puzzling since many satellites had been collecting data, largely uncontested, over sovereign states since the Soviet Union launched the first artificial

Earth satellites in 1957.

The early 1980s was a period when Earth scientists were increasingly thinking about the planet in more holistic and interdisciplinary ways, motivated in part by the new technologies of satellites and computer modeling. This case study reveals what happens when a specific view of

“interdisciplinarity” turns out to be too narrow and the broader reaches of a research program are forgotten. Data concerns were paramount for the political representatives of many developing countries at UNISPACE ‘82, but Global Habitability contained no data policy. Its framers had not considered this crucial detail in the rush of conference preparations. This omission would be

11 Erik M. Conway, Atmospheric Science at NASA: A History (Baltimore: Johns Hopkins University Press, 2008); Erik M. Conway, “Bringing NASA Back to Earth: A Search for Relevance during the Cold War,” in Science and Technology in the Global Cold War, eds. Naomi Oreskes and John Krige (Cambridge, MA: MIT Press, 2014), 251– 72; W. Henry Lambright, “Administrative Entrepreneurship and Space Technology: The Ups and Downs of ‘Mission to Planet Earth,’” Public Administration Review 54, no. 2 (Mar./Apr. 1994): 97–104; W. Henry Lambright, “The Political Construction of Space Satellite Technology,” Science, Technology, and Human Values 19, no. 1 (Winter 1994): 47–69.

95

the primary reason that the initiative was poorly received at UNISPACE. The absence of a

coherent data policy meant that the US could plausibly be left in sole control of massive amounts

of satellite data covering the territories of sovereign states. In a putatively decolonized world

with developing countries issuing calls for NIEO and NWIO, Global Habitability was received

not merely as misaligned with the interests of the developing world, but as antithetical to its

near-term interests in increased economic and political equity. As John Krige suggests, the

transnational movement of information and data is a “social accomplishment” that takes work to

achieve, and so their spread cannot be presumed.12 Global Habitability’s framers appear to have

presumed the neutrality of empirical data and that the availability of more data would benefit

everyone equally.

This case study illuminates the situated, local roots of a proposed global research program that ultimately failed to extend its network of support beyond the US. The proponents

of Global Habitability ignored the need to consider the specific geopolitical realities and

diversity of interests with which they were confronted at UNISPACE ‘82. As a result, scientists

and US representatives never transformed this initiative into an actual research program and,

therefore, the Earth was never studied as a “global habitat.” Scientists and NASA officials

learned important lessons from this experience (chapters three and four). NASA’s next attempt

to construct such a research program—called Earth system science—achieved somewhat more

success.

12 John Krige, “Introduction: Writing the Transnational History of Science and Technology,” in How Knowledge Moves: Writing the Transnational History of Science and Technology, ed. John Krige (Chicago: University of Chicago Press, 2019), 5.

96

THE BIRTH OF GLOBAL HABITABILITY

The exact moments of Global Habitability’s conception are obscured by the fog of time and

memory. What can be discerned of its emergence is Rashomon-like, with not enough evidence

to adjudicate between the various origin stories offered by different historical actors. Shelby

Tilford, NASA’s former Director of Environmental Observations in the Office of Space Science

and Applications (OSSA), claimed that Global Habitability was an obvious and logical extension

of NASA’s targeted ozone depletion research in the 1970s that would rely on satellites with more

comprehensive data collection capabilities. Studying stratospheric ozone was NASA’s first major foray into Earth science research, when a Congressional mandate placed the agency at the head of the US’s ozone research program in 1975.13 According to Tilford, NASA’s ozone

research success supported an expansion of NASA’s Earth science activities into new and

broader areas that eventually encompassed the whole Earth.14 Hans Mark, NASA’s Deputy

Administrator in 1982, traced Global Habitability’s roots to a conversation he had with

atmospheric scientists Richard Goody and Michael McElroy of Harvard University regarding the

need to provide a “philosophical” justification for NASA’s current and upcoming fleet of Earth

observing satellites—like the land-observing Landsat series, the ocean-observing Seasat

successors, and the atmosphere-observing Upper Atmosphere Research Satellite (UARS)—in the

fiscally-constrained Reagan era.15

13 Conway, “Bringing NASA,” 261; W. Henry Lambright, NASA and the Environment: The Case of Ozone Depletion (Washington, DC: NASA, 2005). 14 Shelby Tilford, interview by Rebecca Wright, Washington, DC, 23 Jun 2009, NASA Johnson Space Center, “Earth System Science at 20 Oral History Project,” accessed 5 Jun 2019, https://www.jsc.nasa.gov/history/oral_histories/NASA_HQ/ESS/TilfordSG/tilfordsg.htm. 15 Hans Mark, “Observations of the Earth from Space: A Political Perspective,” 14 Sep 1983, NASA History Office, accessed 12 Feb 2018, https://historydms.hq.nasa.gov/; M. Mitchell Waldrop, “An Inquiry into the State of the Earth,” Science 226, No. 4670 (5 Oct. 1984): 34.

97

Global Habitability-like ideas within NASA trace back at least to 1980 and the work of

NASA’s Advisory Council (NAC). In that year, NAC held a New Directions Symposium to

generate ideas for post-Shuttle activities.16 Among its recommendations was one to study the

Earth as an integrated, rather than fragmented, system—the Earth being then as now the only

known, naturally occurring, closed ecological system—in order to facilitate potential future

space travel and habitation.17 Minutes from a later NAC meeting explicitly stated that the recommendations of NAC’s New Directions Symposium provided, “one stimulus for the more

recent consideration of global habitability programs.”18 Another strand leads back to the passage

of the US National Climate Program Act in 1978. This act prompted NASA to convene a

workshop in 1980 to formulate ideas for a climate observing system that combined existing

meteorological and atmospheric satellite capabilities (like UARS and the Earth Radiation Budget

Experiment) with new platforms that could measure physical variables like ice coverage and

precipitation. The climate observing system never materialized beyond the workshop’s report,

but these ideas were influential on subsequent Earth observing satellite proposals.

All of these threads may have motivated the development of Global Habitability, with

UNISPACE ‘82 providing the proximate catalyst for their consolidation. These threads fit with

what Erik Conway, Neil Maher, and Kim McQuaid have described as NASA’s turn from outer

space to more Earthly activities beginning in the 1970s. With the end of the Apollo program in

1972 and a widespread loss of support for major space “prestige” projects during the Cold War

16 53 participants attended the New Directions Symposium from June 6 to 14, 1980 in Woods Hole, MA. Participants included NASA Administrator Robert Frosch, Freeman Dyson, Luis Alvarez, Verner Suomi, Richard Goody, Lynn Margulis, author James Michener, Michael McElroy, and Berrien Moore. Goody, Margulis, Suomi, and Moore all attended the Global Habitability workshop in June 1982 as well. See: John Naugle, “Report on NASA Advisory Council’s New Directions Symposium, Woods Hole, Massachusetts,” 9 Sep 1980, File 16709, NAC Series, NASA Historical Reference Collection (HRC), Washington, DC. 17 John Naugle, “Report on NASA Advisory Council’s New Directions Symposium, Woods Hole, Massachusetts,” 9 Sep 1980, File 16709, NAC Series, NASA HRC, Washington, DC. 18 “Summary Minutes of the NASA Advisory Council NAC Informal Task Force for the Study of the Mission of NASA, December 1-2, 1982,” 15 Mar 1983, File 16711, NAC Series, NASA HRC, Washington, DC.

98

détente, NASA required new non-Space Race priorities and was more susceptible to public

criticism that called for the agency to undertake activities that could practically benefit the

inhabitants of Earth.19 This turn “back to Earth” meant that conferences focused on the peaceful

uses of space science and technologies—like UNISPACE—were important venues in which

NASA officials could promote these new agency priorities.

Organized by the UN’s Committee on the Peaceful Uses of Outer Space (COPUOS),

UNISPACE I took place in Vienna from August 14 to 27, 1968. 78 states attended, along with nine UN agencies and four other international organizations. It focused on raising the awareness

of countries about the potential benefits of space science and technologies in realms such as

weather forecasting, natural disaster detection, , resource management, and

communications. UNISPACE I was, in essence, a conference meant to advertise, “the importance of practical applications of space exploration and the opportunities available for international co-operation in this field.” Participants viewed space applications expectantly, as

possible solutions to, “the problems of economic and social development[.]”20 All this

optimistic activity took place barely ten years after the Soviets launched the first artificial

satellite into orbit. The Outer Space Treaty of 1967, which insisted on the peaceful use of outer

space, had just been ratified by 108 countries including the two principal space-faring nations,

the US and the USSR. Though there were some exceptions,21 the primary beneficiaries of early

19 Conway, “Bringing NASA”; Neil Maher, Apollo in the Age of Aquarius (Cambridge, MA: Harvard University Press, 2017); Kim McQuaid, “Selling the Space Age: NASA and Earth's Environment, 1958-1990,” Environment and History 12, no. 2 (May 2006): 127-63. 20 United Nations, Committee on the Peaceful Uses of Outer Space (COPUOS), Report of the Committee on the Peaceful Uses of Outer Space (New York: United Nations, 1968), 3. 21 One of NASA’s first major forays into international collaborations came from the agency’s invitation to developing “Third World” countries to build relatively cheap satellite tracking stations to track and collect data from US satellites and spacecraft. These ground stations served a variety of purposes, including the acquisition of meteorological data that could have provided practical weather forecasting benefits to host countries. However, there was no way to actually track data use for these ground stations. See: Ashok Maharaj, “An Overview of

99

space activities were the US, the USSR and a few other developed countries (and certainly not

all citizens of those countries in equal measure).22 Regardless, hopes were high in the 1960s that

outer space—what would be popularized by Gene Roddenberry as the “final frontier”—held

much promise and that its benefits should be spread to all nations. In short, there was a general

optimism about the material benefits to be realized from space activities.

Ten years later, the UN General Assembly agreed to convene a second UNISPACE in

August 1982.23 If UNISPACE I was more informational, UNISPACE ’82 placed greater

emphasis on evaluating how countries, in particular developing countries, actually benefited (or

not) from space activities, and to determine what, if anything, could be done to improve access to

these benefits. While UNISPACE ‘82 has largely faded in memory and has received scant attention from space historians, in 1982 it was a noteworthy event. As NASA’s Tilford relates,

“This was a big deal for space. All of the space agencies and all of the people related to that

from every country in the world that had a space program were going to meet in Vienna[.]”24

Although the conference was a “big deal,” the US formally suspended preparations in 1981

because of a dispute over whether or not an American would become the next Chief of the Outer

Space Division of the UN’s Department of Political and Security Affairs. (Perhaps not

surprisingly, the US said “yes” while the USSR said “no.”) The dispute was resolved when an

American was appointed to the position in January 1982. This allowed the US to resume its

NASA-India Relations,” in NASA and the World: Fifty Years of International Collaboration in Space, eds. John Krige, Angelina Long Callahan, Ashok Maharaj (New York: Palgrave Macmillan, 2013), 216. 22 Reverend Ralph Abernathy’s wagon train protest march to Cape Canaveral for the launch of Apollo 11 on July 16, 1969 and Gil Scott Heron’s “Whitey on the Moon” (1970) serve as forceful reminders that not all segments of US society viewed the space program as equally beneficial or necessary. See: Maher, Apollo, ch. 1. 23 Office of Technology Assessment (OTA), UNISPACE ‘82: A Context for International Cooperation and Competition (Washington, DC: GPO, 1983), 31. 24 Tilford, interview.

100 participation in and preparations for UNISPACE ‘82, but this was a “pyrrhic victory,” according to a report submitted to the US House of Representatives’ Committee on Foreign Affairs in

1983. US preparations for the conference proceeded on a greatly accelerated timeline. Two years of work was crammed into seven months.25 As early as September 1981 and as late as

June 1982, members of Congress raised concerns that this dispute had harmed UNISPACE preparations.26

By March 1982, NASA personnel, in conjunction with external scientists, began weaving together a proposal for a project that would study the Earth as a comprehensive system, what would eventually be called Global Habitability. US officials viewed peaceful initiatives like this as a way to deflect attention away from the issue of the militarization of space that threatened to emerge at the conference.27 The official US position was that the proper forum for space militarization issues was the UN’s Committee on Disarmament, not a conference attended by science and technology experts. The US Office of Technology Assessment (OTA), therefore, recommended that the US develop, “one or more U.S. cooperative initiatives in space” to deflect attention away from the militarization issue.28 This is a concrete example of what Ronald Doel

25 United States House of Representatives, Committee on Foreign Affairs, Report on the Second U.N. Conference on the Peaceful Uses of Outer Space (UNISPACE 1982), August 9–21, 1982, 97th Cong., 2nd Sess. (Washington, DC: GPO, 1983), 20. 26 Letter from Don Fuqua, et al. to James L. Buckley, 21 Sep 1981, File 15692, NASA HRC, Washington, DC. 27 While the Outer Space Treaty (1967) forbade “weapons of mass destruction” from being placed or used in space (“weaponization of space”), it did not, according to US and Soviet interpretations, expressly forbid other kinds of military activities and technologies such as reconnaissance, weather, and navigational satellites or anti-satellite systems funded by militaries or in support of terrestrial military activities (“militarization of space”). In its post- UNISPACE assessment of the pre-conference international context, the US Office of Technology Assessment (OTA) noted that certain recent American actions suggested that the country intended to increase its military presence in space. Soviet “propaganda” further exacerbated the perception of the US as warmongering in space. See: Office of Technology Assessment, Congress of the United States, “UNISPACE ‘82: A Background Report,” 14 Jul 1982, File 15692, NASA HRC, Washington, DC; OTA, UNISPACE ‘82, 28. 28 Office of Technology Assessment, Congress of the United States, “UNISPACE ‘82: A Background Report,” 14 Jul 1982, File 15692, NASA HRC, Washington, DC.

101

calls, “deliberate efforts by governments to utilize science and scientists to reach foreign policy

goals.”29 In this case, the foreign policy goal was evasion rather than invasion.30

With UNISPACE ‘82 fast approaching, NASA officials worked quickly to develop a

sketch of Global Habitability. Tilford wrote to NASA upper management on March 12 to inform

them that, as they had directed, Tilford and atmospheric scientists Richard Goody and Michael

McElroy had held, “several discussions regarding an ‘Earth Systems’ initiative” and that they

would conduct a follow up summer workshop to, “prepare documentation, plans, etc., for such

an initiative.”31 On March 25, Hans Mark wrote to NASA Administrator James Beggs about

UNISPACE ‘82 preparations. Mark noted the work being done by Goody on a potential

UNISPACE initiative for a, “world-wide observational program to deal with small long-term changes in the oceans and atmosphere that may have adverse consequences if they are not well understood.” Mark threw his support behind this proposal, noting that it, “would be both technically feasible and politically attractive.”32 It was politically attractive because, for Mark

and other NASA officials, it would be a wholly peaceful and environmental initiative that the US

could use at UNISPACE to deflect from questions about space militarization. Indeed, Global

Habitability was so attractive that Kenneth Pedersen—NASA’s Director of International

Affairs—presumed that its successful introduction at UNISPACE would pave the way to

29 Ronald E. Doel, “Scientists as Policymakers, Advisors, and Intelligence Agents: Linking Contemporary Diplomatic History with the History of Contemporary Science,” in The Historiography of Science and Technology, ed. Thomas Söderqvist (Amsterdam: Harwood Academic, 1997), 216. 30 Using science to achieve geopolitical goals fits within the US government's broader use of “cultural diplomacy” during the Cold War in its efforts to fight against and contain the spread of communism. See: Audra J. Wolfe, Freedom's Laboratory: The Cold War Struggle For the Soul of Science (Baltimore, MD: Johns Hopkins University Press, 2018). 31 Shelby Tilford to NASA Deputy Associate Administrator and Associate Administrator for OSSA, 12 Mar 1982, Global Habitability File, Box 18045, NASA HRC, Washington, DC. 32 Hans Mark to Administrator, 25 Mar 1982, Global Habitability File, Box 18045, NASA HRC, Washington, DC.

102 increased domestic political support with the US Office of Management and Budget (OMB).33

The OMB was the first stop for US agencies looking to receive discretionary funding from

Congress.34

At Woods Hole, MA from June 21 to 26, 49 NASA personnel and external scientists gathered for the first—and, as it would turn out, only—workshop on Global Habitability. Goody coordinated the proceedings and served as chairman. Many prominent scientists attended, notably Lynn Margulis, James Hansen, and Paul Crutzen. Future Earth System Sciences

Committee (ESSC) members—James Baker, Daniel Botkin, Moustafa Chahine, Berrien Moore,

Ronald Prinn, and Wilford Weeks—also participated. (Though future ESSC chair Francis

Bretherton did not attend the final workshop, he was one of 28 individuals who participated in a pre-workshop steering committee meeting in April 1982.35) The scientists at the Global

Habitability workshop represented a wide swath of Earth and environmental science disciplines: meteorology, biology, ecology, atmospheric chemistry and physics, oceanography, hydrology, glaciology, geology, and geophysics. Nearly all were from Europe and the US; non-Western, developing countries were notably unrepresented. On July 7, barely a month before the start of

UNISPACE ‘82, Goody submitted the workshop’s final report, Global Change: Impacts on

Habitability.36

33 Kenneth Pedersen to Associate Administrator of Space Science and Applications, 1 Jul 1982, Global Habitability File, Box 18045, NASA HRC, Washington, DC. 34 In the US, budget proposals for discretionary funds are submitted by the various government agencies to the Office of Management and Budget, part of the Executive Office of the President. The Office of Management and Budget (OMB) and the President then table a proposed budget to submit to Congress. Congress authorizes all national expenditures and final appropriations, either agreeing with the President’s proposed budget or, not infrequently, rejecting part(s) of it. These can be contentious and drawn out processes that occasionally result in partial government shutdowns if agreements are not reached, like the most recent shutdowns during the Trump Administration in 2018 and 2018-2019. 35 Meeting request from Shelby G. Tilford, 8 Apr 1982, Global Habitability File, Box 18045, NASA HRC, Washington, DC. 36 JPL, Global Change.

103

At a mere 13 pages, this report was clearly not intended to provide concrete and specific

details about the new NASA initiative. Prepared hurriedly in the two months prior to the start of

UNISPACE ‘82 on August 9, its purpose was to briefly answer three questions. Was Global

Habitability, broadly speaking, feasible? Was NASA the proper agency to lead it? Should

Global Habitability be unveiled at UNISPACE ‘82 as a new US initiative? The report’s answer

to all three questions was “yes,” perhaps not surprisingly given the political agenda and tight

timeline. To the question why NASA and not some other agency, the report answered that, “The

short answer is that NASA can do it and no other Federal agency can.” NASA was, according to

the report, well-suited to lead this global initiative given its technical space capabilities and its,

“experience with interdisciplinary project management of the kind required for this program.” 37

With respect to the presentation of Global Habitability as a US initiative at UNISPACE ‘82, the

report stated: “While we are conscious of the great difficulties in effective execution, we

nevertheless believe that the step is appropriate for reasons both of feasible science and of good

international policy.”38 Though the report noted that “effective execution” would be difficult, international acceptance was, by and large, implicitly and unequivocally assumed.

Global Habitability lived up to the “global” part of its name. The report called for a comprehensive study of the whole planet to better understand the interacting physical, chemical, and biological processes at work in the air, land, ice, and water on decadal timescales, the timescales most relevant for life. On these timescales, the Earth’s oceans, atmosphere, land, cryosphere, and biosphere, “operate as a coupled system not only in their physical interactions, but also through chemical and biological processes.”39 This was, “a fascinating but exasperating

37 Project Viking and the Upper Atmospheric Research Program were two examples given of past successful interdisciplinary projects. See: JPL, Global Change, 2. 38 JPL, Global Change, 2. 39 Ibid, 3.

104

symbiosis.”40 Scientists, however, now possessed better tools to collect measurements and

model these interconnections, which would help reduce exasperation: “through satellites,

computers, and modem communications humanity has the scientific tools to carry out the

required research.”41 The report noted the need for interdisciplinary collaboration to better

understand the interconnecting processes affecting living conditions on the planet: “Physics

cannot be separated from chemistry and biology by considering them consecutively. Nor can

problems in the ocean be treated in isolation from those on the land or in the atmosphere. The

necessary [scientific] questions relate to the interaction of land, sea, and atmosphere.”42

Global Habitability also lived up to the “habitable” part of its name. With the proposed

initiative, scientists would work towards understanding how Earth processes were changing both

naturally and through human actions, and what these changes might entail for the planet’s future

habitability. For the workshop participants, habitability meant the ability of Earth to support communities of plants and animals with the requisite air, water, and nutrients, as well as favourable climatic conditions. The larger aim was to understand what made Earth uniquely

habitable, which is to say unique when compared to the biological barrenness of Venus and Mars

revealed by the Mariner and Viking exploration missions in the 1960s and 1970s. (This was

predicted by James Lovelock based on the planets’ atmospheric composition -- see chapter

one.)43 Further, the report gave urgency to the research by emphasizing the environmental

problems arising from the ways in which human activities were altering the planet’s

environment. Increasing levels of atmospheric carbon dioxide, stratospheric ozone depletion,

acid rain, air and water pollution, desertification, deforestation, overgrazing, and the depletion of

40 Ibid, 1. 41 Ibid, 4. 42 Ibid. 43 See: Conway, Atmospheric Science at NASA, 94.

105

freshwater all potentially threatened the Earth’s future habitability. The choice of timescale for

the initiative favoured certain Earth science disciplines over those focused on longer timescales

(for instance geology and geophysics, which often focus on long-term processes like those in the

Earth’s core and mantle), but this was not much of an issue since Global Habitability was only

broad outline for a potential future program. It was not concrete enough to actually threaten or

marginalize research. The question of timescale would become a much greater issue when the

ESSC constructed Earth system science, a more detailed Earth science research program with a

concrete implementation strategy (chapter three).

Global Habitability would use satellites to produce, and computer models to analyze, global datasets, integrating ever more variables into these models. This meant greater interdisciplinary collaboration. The workshop attendees summarized the situation as follows:

“As the knowledge in each discipline has grown, so have the boundaries of investigation. We have now reached the point where the boundaries of each discipline are overlapping, and the next

step forward can only come from an interdisciplinary research program. Such a program

requires dedicated scientists, sophisticated tools using advanced instruments, computer and

satellite technology, and strong managerial leadership.”44 Priority was given to the interactions

between the Earth’s major components, rather than studying parts in disciplinary isolation. The

report recommended no specific program, but offered assorted potential research foci: the

workings of the hydrological cycle in the atmosphere, oceans, and on land; the dynamics and

composition of the upper layers of the oceans; coupling between the atmosphere and land; the climatic role of snow and sea ice; nutrient and toxic compound cycling between the atmosphere,

land, rivers, coastal zones, and deep oceans; physical and chemical processes in the atmosphere

44 JPL, Global Change, 4.

106

and oceans; and determining the total biomass of the Earth’s surface and its decadal rate of

change with increased accuracy. All potential foci centered around dynamic and complex

interactions, and would require collaboration between scientists from the different Earth

sciences. To study, for instance, the coupling between the atmosphere and land would require

observations of surface properties like albedo, heat capacity, soil moisture, vegetation cover,

evapotranspiration rates, and roughness, along with atmospheric water content, clouds,

precipitation, temperature, and wind speeds, and so might require expertise from ecologists,

biologists, hydrologists, geologists, geophysicists, atmospheric physicists, and atmospheric

chemists. Observational data collected from satellites, aircraft, balloons, and ground stations

would be combined with the results of laboratory experiments and theoretical work to develop

several comprehensive models to provide fundamental knowledge of the planet and to help in

policy decision making.45

This proposed research program traversed national as well as disciplinary boundaries and

so would require international cooperation for the collection in-situ observations. The report explicitly emphasized: “A global habitability program must be international in scope and requires the cooperation of the international science community. No one nation could carry it out alone. These considerations are recognized by scientists and policymakers around the world.”46 For this reason, the report recommended that NASA officials encourage participation

from all countries and that Global Habitability research contribute to ongoing and planned

international scientific endeavours like those conducted by the UN Environment Programme

(UNEP), the World Meteorological Organization (WMO), the World Climate Research

Programme (WCRP), the International Council of Scientific Unions (ICSU), and the UN Food

45 Ibid, 5-7. 46 Ibid, 13.

107 and Agriculture Organization (FAO).47 Importantly, the report did not recommend that NASA solicit input from individual countries about the framework for this research, or the implications of data collection and analysis on such a vast scale.

Given the large-scale, global nature of many of the processes to be studied—and the fact that it was to be a NASA-led initiative—Earth observing satellites were to play a crucial role in the Global Habitability initiative. Building on the earlier proposed climate observing system, the report split satellite program elements into three categories. The first category included current platforms like weather satellites that had been collecting meteorological data since TIROS 1’s launch on April 1, 1960.48 Land observing satellites like the Landsat series complemented these extant meteorological observing capabilities.49 Secondly, the report identified satellites that were “ready for execution” and in need of immediate funding, listing two specific examples:

UARS and the Ocean Topography Experiment (TOPEX). Plans for UARS incorporated a radiometer to sense chlorine monoxide, ozone, and water vapour, as well as two infrared spectrometers to sense chemical species in the stratosphere like chlorine, nitrogen, and fluorine.50 TOPEX would augment the observations made by the short-lived but highly successful Seasat, the first ocean-dedicated satellite that flew for 106 days in 1978.51 TOPEX’s

47 Ibid, 13. 48 Conway, Atmospheric Science at NASA, 28-9. 49 Landsat 1 launched on 23 July 1972 with two remote sensors: a return beam vidicon (a type of television camera with a very high resolution), and a multispectral scanner (which collected radiometric data from the Earth from four spectral bands). With almost each new satellite in the Landsat program, the mounted instruments grew increasingly more sophisticated, with such additions as new bands (e.g. to detect thermal infrared radiation) for the multispectral scanner and a thematic mapper (a higher-resolution multispectral scanner collecting data from seven spectral bands). Subsequent satellites were launched in 1975, 1978, 1982, 1984, 1993, 1999, and 2013. Most recently, Landsat 8 was launched on 11 February 2013, with Landsat 9 slated to launch in 2020. See: Pamela E. Mack, Viewing the Earth: The Social Construction of the Landsat Satellite System (Cambridge, MA: MIT Press, 1990), 68-73; “Landsat Science: History,” NASA, accessed 23 April 2019, http://landsat.gsfc.nasa.gov/?page_id=2281. 50 Conway, Atmospheric Science at NASA, 146-7. 51 If an electrical short out had not prematurely ended the mission, Seasat would have collected tens of thousands of oceanographic measurements. Seasat carried five instruments collecting data on variables like sea surface height, wind speed and direction, sea surface temperature, and sea ice conditions. See: Erik M. Conway, “Drowning in

108

radar altimeter would collect continuous sea-surface elevation observations to increase knowledge of the movement of heat and chemical species in the oceans.52 Lastly, entirely new

satellites and instrumentation would be required to collect data on poorly understood or

inadequately observed variables like hydrological and biogeochemical cycles in the troposphere,

sea ice topography and extent, soil moisture, and evapotranspiration.53 The report framed

satellite data collection as central to the project by making in-situ measurements and laboratory experiments merely complementary to satellite work. It may then be somewhat surprising that the report provided no details about how satellite data would be collected or stored, or who would have access. Instead, the report emphasized the difficulties involved in effectively analyzing global datasets. It identified the scientific usability of satellite data as a “high priority.”54 By doing so, the report implicitly assumed the neutrality and beneficence of

scientific data, and that unhindered data access was an unquestioned good. The omission of a

robust data policy would significantly contribute to the rejection of Global Habitability at

UNISPACE ‘82.

NEW WORLD ORDERS

From August 9 to 21, delegates from 94 member states along with 45 intergovernmental and

non-governmental organizations met in Vienna for UNISPACE ‘82. They attended plenary

talks, smaller committee deliberations, poster sessions, and technical demonstrations. They

Data: Satellite Oceanography and Information Overload in the Earth Sciences,” Historical Studies in the Physical and Biological Sciences 37, no. 1 (2006): 127-51. 52 Conway, “Drowning in Data,” 145. 53 JPL, Global Change, 5-12; Conway, Atmospheric Science at NASA, 220-3. 54 JPL, Global Change, 12.

109 enjoyed evening lectures from distinguished space luminaries like Arthur C. Clarke and Carl

Sagan. They toured the exhibition hall where delegations presented exciting displays of the latest space science and technology.55 The US delegation consisted of representatives from the

State Department and NASA, the American novelist James Michener—who spoke at an

American gala on the “virtues of the space shuttle”56—as well as current and future astronauts.57

Throughout, US delegates worked hard to promote Global Habitability. NASA Administrator and head of the US delegation, James Beggs, publicly introduced the project in his opening

UNISPACE speech, declaring that, “increased scientific understanding of environmental problems and improved methods of forecasting are needed if we are to enhance our ability to address issues relating to overall global habitability in an effective and efficient manner.”58

Tilford and Goody presented more details on the initiative in a subsequent poster session.59

According to one reporter, the US even, “devoted a fair chunk of an official American gala evening...to talk about what it [Global Habitability] means.”60

Focused on the positive features of the initiative—its peacefulness, environmental focus, and supposedly internationally-inclusive character—NASA officials anticipated that Global

Habitability would be well-received at UNISPACE. In its pre-conference report, the OTA suggested that such an initiative, “could be a welcome plan to many other countries of the world, particularly the developing countries[.]” As a result, it, “could give the United States a major

55 United Nations, Report on the Second United Nations Conference on the Exploration and Peaceful Uses of Outer Space (United Nations, 31 Aug 1982), 142-51. The US’s space exhibit was titled “Working Together to Benefit Spaceship Earth.” See: United States House of Representatives, Committee on Foreign Affairs, Report, 22. 56 Peter Marsh, “Only the Rich Become Richer in Space,” New Scientist (19 Aug. 1982): 473. 57 United States House of Representatives, Committee on Foreign Affairs, Report, 10. 58 James M. Beggs, “Ambassador Beggs’ Statement, Aug. 10, 1982,” Department of State Bulletin 83, no. 2071 (Feb. 1983): 70. 59 United Nations, Report, 146. 60 Marsh, “Only the Rich,” 473.

110

political advantage at UNISPACE ‘82.”61 The optimism is perhaps understandable, but it would turn out to be unfounded. This peaceful, environmental program was not warmly welcomed by other countries. In fact, NASA personnel at UNISPACE ‘82 reported a decidedly unfavourable reception. Tilford, who presented the initiative with Goody during a poster session, described vivid memories of the rejection: “We were shot down!...No one accepted it. They would not buy into it. They thought it was the United States trying to take over the world, and that we were going to keep all of the data, and we were going to have all of the information on their countries, and they didn’t want that.”62 As Conway describes it, the initiative received, “near-universal condemnation.”63 He notes that Tilford characterized the response at UNISPACE to Global

Habitability as, “not merely disinterested” but “openly hostile[.]”64

After UNISPACE, the words “Global Habitability” were banned from NASA budget

letters because, according to NASA’s Deputy Administrator Hans Mark, “we know it triggers

negative reactions.”65 In Science, journalist M. Mitchell Waldrop suggested that the “reviews”

of the initiative in Vienna were “scathing.” According to Waldrop, “Somehow the message

came across as ‘Here’s what NASA’s going to do. Join us.’” Many delegates, particularly those

from the developing world, “were insulted by the implied condescension[.]”66 Mark recounted,

with the benefit of hindsight, that international initiatives must take into account the various

“sensitivities” of other countries, perhaps not quite grasping the underlying issue that other

countries had genuine concerns, not just delicate “sensitivities.”67 In their development of

61 Office of Technology Assessment, Congress of the United States, “UNISPACE ‘82: A Background Report,” 14 Jul 1982, File 15692, NASA HRC, Washington, DC [emphasis added]. 62 Tilford, interview. 63 Conway, “Bringing NASA,” 260. 64 Conway, Atmospheric Science at NASA, 224. 65 Mark, “Observations,” 20. 66 Waldrop, “An Inquiry,” 34. 67 Mark, “Observations,” 20.

111

Global Habitability, NASA officials and affiliated scientists failed to grasp these concerns. The

seriousness with which US Congressional representatives viewed the conference outcome is

demonstrated by the fact that the House Committee on Science and Technology and the Joint

Economic Committee ordered a review of any, “potential adverse effects that UNISPACE ‘82

might have on US interests in outer space[.]”68 The OTA also prepared a post-conference analysis of the US’s dismal performance at UNISPACE ‘82. Years later, ESSC program officer

Laura Lee McCauley remarked to ESSC colleagues that, “I have heard horror stories about the global habitability presentation at UNISPACE a few years back.”69 As late as 1986, four years

after the event, ESSC members reported that they were, “still stinging from global

habitability[.]”70

To understand the source and nature of this international backlash, it is important to

examine the broader political and economic contexts of the 1970s and early 1980s in which

UNISPACE ‘82 originated. The UN General Assembly decision to hold another space

conference was itself a manifestation of changing international dynamics. Beginning in the

1970s and continuing into the early 1980s, developing countries experienced new power on the

world stage.71 The 1960s was a time of tremendous political upheaval in many parts of the world still fighting for independence from colonial rule or dealing with post-revolutionary internal conflicts. However, by the 1970s many of these new nations had achieved relative

68 OTA, UNISPACE ‘82, iii; Conway, Atmospheric Science at NASA, 224. 69 Laura Lee McCauley, “Update on ESSC [Earth System Sciences Committee] Activities and Plans, [1986], Folder 71, Box 3, Earth System Sciences Collection, National Center for Atmospheric Research (NCAR), Boulder, CO. 70 ESSC [Earth System Sciences Committee] Conference Call Notes, [1986], Folder 100, Box 4, Earth System Sciences Collection, NCAR, Boulder, CO. 71 This paper uses the “developed/developing” language employed by historical actors at UNISPACE ‘82, as opposed to other languages of distinction like “North/South” or “First/Second/Third World.” In the post-World War II “development” language, the rich, industrialized, developed countries of the world served as models for the poor, agrarian, less industrialized ones. Recent scholarship has problematized and historicized this “development discourse” by showing the conditions of its creation and the often-detrimental impacts it has had on those so-called “developing” countries. See: Arturo Escobar, Encountering Development: The Making and Unmaking of the Third World (Princeton, NJ: Princeton University Press, 2011 [1995]).

112

political stability, and so turned their focus to economic and social development.72 Following

decolonization, developing countries enjoyed numerical superiority over developed countries.

This numerical superiority could, and often did, translate into control of the UN agenda in most

forums, apart from the Security Council.73 There were a number of ways that this new-found power played out, but with respect to Global Habitability and UNISPACE ‘82, the most salient were the restructuring calls associated with the New International Economic Order (NIEO) and the New World Information Order (NWIO) that provided part of the rationale for holding

UNISPACE ‘82 and contributed to the backlash against Global Habitability.

In an effort to better coordinate and articulate their activities and positions on the world stage, 77 developing countries joined together in 1964 to form what was later called the Group of

77. This group—whose membership has continued to grow since its formation and now numbers 134—turned their common interests into political leverage by negotiating and voting as a coordinated bloc to promote developing world interests at the UN. This included the introduction of a resolution calling for the “Establishment of a New International Economic

Order” (NIEO) that was passed by the UN General Assembly on May 1, 1974.74 The NIEO

called for structural modifications to the that would result in more equitable

economic relations and more sovereign control over activities within national borders to, “correct

72 Independence struggles did not end with the 1960s. Many territories—including Guinea-Bissau, Angola, Mozambique, and Zimbabwe —fought for independence throughout the 1970s and beyond. Today, the UN's Special Committee on Decolonization maintains a list of “non-self-governing territories.” See: “Non-Self- Governing Territories,” United Nations, accessed 10 Mar 2020, https://www.un.org/dppa/decolonization/en/nsgt. 73 Decolonization led to an increased number of states, with the rise from 51 original UN states in 1945 to 110 by 1962 largely attributable to this process. Sovereign equality in the UN meant that in almost all UN forums every country’s vote counted equally, and therefore the more numerous developing countries could outnumber other countries and influence the UN agenda. One major exception here is the UN Security Council, where five countries (US, UK, France, Russia/Soviet Union, and China) have veto power. See: Stephen D. Krasner, Structural Conflict: The Third World Against Global Liberalism (Berkeley, CA: University of California Press, 1985), 7-10. 74 “The Group of 77 at the United Nations,” accessed 11 June 2018, http://www.g77.org/doc/index.html; Daniel J. Sargent, “North/South: The United States Responds to the New International Economic Order,” Humanity 6, No. 1 (Spring 2015): 201.

113

inequalities and redress existing injustices” that resulted from the legacy of colonialism.75

Though such calls rarely materialized into concrete changes, the NIEO concept did provide the

rationale for a number of international actions, including the idea to hold a second UNISPACE

conference in 1982. This idea was first raised in a UN organization in 1974 (by COPUOS’s

Scientific and Technical Subcommittee), the same year that the UN General Assembly passed

the NIEO declaration, suggesting an overlap between Earthly appeals for economic and social

equity and outer space.76 Indeed, the final report for UNISPACE ‘82 noted that a number of

country delegates, “saw the [UNISPACE] Conference in the context of the ongoing efforts to

promote the new international economic order which space technology, if properly used, could

support.”77 Space technologies, like Earth observing satellites, offered the promise of data that

might allow sovereign states to act more autonomously in the management of their resources and

to better prepare for environmental disasters like droughts, floods, fires, and pest infestations.

Calls from the developing world for more equitable economic relations came with calls for more equitable access to and flow of information. These concerns were linked to broader calls for what was referred to in the 1970s and 1980s as a “New World Information Order”

(NWIO), articulated most forcefully in UNESCO’s MacBride Report (1980).78 This report detailed what Joseph Nye would much later describe as “soft power.”79 That is, the

asymmetrical communication order in which rich, industrialized countries in the North

75 United Nations, General Assembly, Declaration on the Establishment of a New International Economic Order, A/RES/S-6/3201 (1 May 1974). 76 United Nations, Report, 24. 77 Ibid, 123 [emphasis added]. 78 MacBride Commission, Many Voices, One World: Towards a New, More Just and More Efficient World Information and Communication Order (New York: Unipub, 1980). The “New World Information Order” was also called the “New World Communication Order” or the “New World Communication and Information Order.” For consistency, this paper uses the singular phrase “New World Information Order” to refer to all these nomenclature variations. 79 Joseph Nye, “Soft Power and American Foreign Policy,” Political Science Quarterly 119 (2004): 255-70.

114

dominated the one-way flows of mass media and other cultural products to the poorer,

developing countries of the South.80 A legacy of colonial relations, these asymmetries carried over into unequal access to other kinds of information—including data from Earth observing satellites—due to economic or technical barriers that prevented developing countries from obtaining access. According to the MacBride Report, scientific and technical data, “is today a vital economic resource…[that] ought to be more generously and widely shared[.]”81 At the

same time, countries, particularly those in the developing world, demanded greater control over

who could access data on their territories. The MacBride Report noted that, “A major concern in

this area is linked with the sovereignty of countries surveyed. Technical facilities are so

powerful that important data about a developing country may now be better known in some

foreign capitals than by the national government.”82

At stake was not just the collection but also the dissemination and utilization of data.

Earth observing satellite data on one sovereign nation might be used by another sovereign nation

(or, just as likely, a transnational corporation), and there was an absence of international laws

and bilateral agreements to regulate these information flows.83 Earth observing satellites had

been collecting data on countries since the first US and USSR reconnaissance satellites launched

in the early 1960s.84 Civilian meteorological satellites like TIROS had been in operation since

1960. But only with the launch of Landsat-1 in 1972 was satellite data more widely collected

80 As an example, James Brennan notes that, in the 1960s and 1970s, over 90 percent of print and broadcast news around the world was controlled by four agencies: Reuters (UK), Agence France-Presse (France), Associated Press (US), and United Press International (US). See: James R. Brennan, “The Cold War Battle Over Global News in East Africa: Decolonization, the Free Flow of Information, and the Media Business, 1960–1980,” Journal of Global History 10 (2015): 341. 81 MacBride Commission, Many Voices, 25. 82 Ibid, 79. 83 COPUOS, Report of the Legal Sub-Committee on the Work of Its Twenty-First Session (1-19 February 1982), 24 Feb 1982, File 15692, NASA HRC, Washington, DC. 84John Cloud, “Imagining the World in a Barrel: CORONA and the Clandestine Convergence of the Earth Sciences,” Social Studies of Science 31, no. 2 (Apr. 2001): 231-51; Mack, Viewing the Earth.

115

and available.85 Landsat data dissemination occurred on a “nondiscriminatory basis,” meaning

that it was equally available to all users for purchase.86 Even so, equal availability of satellite data was not the same as equal capacity to purchase, process, and utilize that data. Those possessing the requisite computing capacity and expertise for data processing had a clear advantage. Neil Maher outlines NASA efforts to promote international use of Landsat data via training programs for foreign scientists and engineers. Yet, as Maher emphasizes, NASA

maintained ultimate authority over the satellite, ground station, and data analysis hardware, and it

further determined which countries could participate in the Landsat program and the kinds of

research foreign scientists could undertake with NASA funding.87 According to Maher,

asymmetries in scientific knowledge, technological capacity, and political power remained ever

present.88 NASA control over access to and use of Landsat data replicated the asymmetrical

power relations of colonialism in all but name. What Landsat showed was that, even if in theory

everyone was free to use satellite data, in practice some were more free than others.

As Earth observing satellites became more technically sophisticated, numerous, and

widely used, thorny legal issues emerged. These concerns were not just “sensitivities” raised

outside the US but were acknowledged by American jurists too. In 1977, American legal scholar

Hamilton DeSaussure charted the many nuanced concerns related to, “the ever increasing

perceptiveness of the orbiting eye in space.” In a presentation to the New York Bar Association

he gave a lengthy outline of the many difficult legal questions raised by space observation:

85 Mack, Viewing the Earth. 86 Office of Technology Assessment (OTA), Remote Sensing and the Private Sector: Issues for Discussion (Washington, DC: GPO, 1984), 7. 87 Neil M. Maher, “Bringing the Environment Back In: A Transnational History of Landsat,” in How Knowledge Moves: Writing the Transnational History of Science and Technology, ed. John Krige (Chicago: University of Chicago Press, 2019), 207-14. 88 Ibid, 204.

116

How is this activity [remote sensing data collection] presently governed by the Outer Space Treaty and by general international law? What should be the defined limits for remote sensing activity? Is sovereignty infringed by the sensing and distribution of data about resources within the subjacent state without the permission of that state? Do states have a right to forbid or control the sensing of their territory? Do they have a right to object to the distribution of data about their territory to third states? Do they have a prior right to receive and analyze this data prior to distribution to a third party? Is the data generated by remote sensing satellite or the information derived from this data subject to individual proprietary rights? Does remote sensing constitute an invasion of privacy, either at the sovereign or individual level? Should special rules be applicable to remote sensing in national border areas or nationally declared restricted zones? Do ground receiving stations which collet [sic] data about the territory of adjacent countries have any special obligations toward those neighbors? Should data access be placed under international control? Should there be one central international data bank for the collection, storage, and processing of all acquired data? Is a multilateral treaty necessary or desirable to provide for the regulation of remote sensing activity?89

DeSaussure wanted to emphasize that many countries viewed satellite data as a potential

economic threat, since this data could enable economic exploitation by other parties—including

transnational corporations—of a sovereign state’s agriculture and natural resources.90 These

concerns may have been justified given that a 1976 NASA-sponsored survey of several hundred users of Earth observing satellite data showed that “industry” was the largest user group. In an

effort to codify protections against economic exploitation of data, Argentina, Chile, Venezuela,

and Mexico presented draft satellite data regulations to COPUOS’s Legal Subcommittee in the

early 1970s that would prohibit the collection of satellite data on a sovereign state’s natural

resources without prior consent.91 While it did not adopt these specific draft regulations, the

Legal Subcommittee undertook a study of the issue. Pamela Mack notes that the subcommittee

89 Hamilton DeSaussure, “Remote Sensing by Satellite: What Future for an International Regime?” The American Journal of International Law 71, no. 4 (Oct. 1977): 719. 90 For instance, satellite data had already been used to identify new copper ore deposits in Pakistan. It might also be used to identify surplus or deficient agricultural commodity production, knowledge that could be economically exploited. See: DeSaussure, “Remote Sensing by Satellite,” 714. 91 DeSaussure, “Remote Sensing by Satellite,” 714-5, 720.

117

did reach an agreement on a number of issues, but it did not resolve the main issue relating to the

legality of one country using a space-based platform to collect data on another country, or the legality of disseminating this information to third parties.92 The COPUOS Legal

Subcommittee’s report of its February 1982 meeting—its last before UNISPACE ‘82 in

August—outlined the points of disagreement, which largely centered around the precise

permissions that would (or would not) be required for “third party” access to remotely sensed

data on sovereign states.93

To understand the fate of Global Habitability, it is important to note that, in 1982, the

international legal issues regarding satellite data collection and dissemination were contentious

and far from settled. At issue were both economic structural inequalities and national

sovereignty. Developing countries desired to reap any benefits of satellite data but they also

demanded protections against foreign exploitation and invasive foreign surveillance. This issue

would gradually fade throughout the 1980s as the benefits of such data came to be viewed as

outweighing potential harms, but in 1982 it was still a sensitive issue.94 UNISPACE ‘82

materialized in the political and economic contexts of NIEO and NWIO, where outer space

became another contested territory of colonial expansion. Developing countries sought a share

of any benefits that might accrue from the exploitation of space. The first calls to hold another

UNISPACE occurred in tandem with the UN General Assembly's passing of the NIEO

declaration in 1974. According to the OTA, the economic needs of developing countries formed

92 Mack, Viewing the Earth, 187. 93 COPUOS, Report of the Legal Sub-Committee on the Work of Its Twenty-First Session (1-19 February 1982), 24 Feb 1982, File 15692, NASA HRC, Washington, DC. 94 This concern was high in 1982, in part, due to Landsat’s imminent move to operational status, and its potential transfer to private ownership that threatened to increase data access prices. See: Mack, Viewing the Earth, 188. The Landsat satellite system was transferred from public to private control in 1984. However, data continuity and foreign competition pressures led to Landsat’s transfer back to full governmental control in 1992. See: Ray A. Williamson, “The Landsat Legacy: Remote Sensing Policy and the Development of Commercial Remote Sensing,” Photogrammetric Engineering & Remote Sensing 63, no. 7 (Jul. 1997): 877-85.

118

the “raison d’etre” of UNISPACE ‘82: “The developing countries want a larger share of the

world economy and view access to space technology as necessary to accelerate development and

achieve greater parity with industrialized nations.”95 The OTA report continued: “The

UNISPACE ‘82 agenda is to prepare a report for the U.N. General Assembly stating the present and projected state of space technology and its applications, with emphasis on the developing countries….As such it will serve as a statement of the interests and aspirations of the developing world with respect to space science and technology.”96 Developing countries themselves had

provided the driving force for holding another UNISPACE conference, with the US and USSR

agreeing to participate only reluctantly.97

EVIDENCE OF ABSENCE

By unilaterally developing Global Habitability without an adequately articulated data policy,

NASA personnel and US representatives demonstrated a blindness to the surveillance

imperatives of their proposed program, the existing inequalities in the distribution and processing

capabilities of satellite data, as well as a lack of awareness about the importance of the

international geopolitical context. The extent of this blindness is evidenced by NASA officials’

surprise at being confronted by a hostile international response to the project. Prior to and during

the conference, NASA personnel and US representatives failed to recognize UNISPACE ‘82—

and the UN in general—as being, to borrow from Mary Louise Pratt, a postcolonial contact zone.

Pratt describes contact zones as, “social spaces where disparate cultures meet, clash, and grapple

95 Office of Technology Assessment, Congress of the United States, “UNISPACE ‘82: A Background Report,” 14 Jul 1982, File 15692, NASA HRC, Washington, DC. 96 Ibid [emphasis added]. 97 United States House of Representatives, Committee on Foreign Affairs, Report, 20.

119

with each other.”98 In this case the clash was over outer space technologies and a proposed Earth science research program with global reach. The UNISPACE ‘82 postcolonial contact zone was still suffused with a bitter legacy of asymmetrical colonial power relations. Developing countries, historically marginalized on the world stage, had just begun to exert more political power in international settings like the UN through coordinated actions in promotion of common interests. UNISPACE ‘82 was a place where a developed country like the US confronted resistances from developing countries like those in the Group of 77. It was a social space where delegations from developed countries should no longer have expected to simply impose an agenda without contest on docile developing countries without prior consultation and negotiation. With NIEO and NWIO concerns at a peak, developing countries were not simply being “sensitive.” They expected consultations and negotiations for global programs, as well as a recognition of their concerns. US representatives’ lack of sensitivity may have been a consequence of haste or ignorance, but either way, Global Habitability was a global data collection proposal clearly out of step with the concerns of developing countries in 1982.

As might have been anticipated, the most prominent and contested issue at UNISPACE

‘82 was satellite data collection and dissemination.99 Conference delegates from the developing world expressed the double-edged character of satellite remote sensing data. Brazilian President

H.E. Joao Baptista de Oliveira Figueiredo highlighted the “marvellous” benefits that might come

from more extensive knowledge of natural resources, vegetation, agriculture, weather, and

pollution, while at the same time cautioning that this knowledge was “dangerous.”100 Mexico,

98 Mary Louise Pratt, Imperial Eyes: Travel Writing and Transculturation (New York: Routledge, 1992), 4. 99 UNISPACE ‘82 background documents like the COPUOS Legal Subcommittee report provide additional evidence supporting the contentiousness of this issue, explicitly noting the central importance of remote sensing data policies for developing countries at UNISPACE ‘82. See: COPUOS, Report of the Legal Sub-Committee on the Work of Its Twenty-First Session (1-19 February 1982), 24 Feb 1982, File 15692, NASA HRC, Washington, DC. 100 United Nations, Report, 156.

120

on behalf of the Group of 77, submitted a proposal at the conference regarding these remote

sensing issues, which asserted the need to respect national sovereignty by obtaining prior consent

for the collection and dissemination of satellite data.101 A Chinese UNISPACE ‘82

commemorative stamp forcefully captured this data ambivalence (Figure 2.1).102 The stamp

listed the full name of the conference (The Second United Nations Conference on the

Exploration and Peaceful Uses of Outer Space) superimposed on an oversized predator-like

satellite flying above the Earth.103 The satellite’s detector beams were trained on the entire

visible planetary surface, which was represented as an abstract sphere traversed by lines of

longitude and latitude. The stamp juxtaposed the rhetoric of the peaceful use of scientific

knowledge obtained via satellites with techno-optimistic (or techno-pessimistic, depending on

one’s disposition) notions of total surveillance and, implicitly, the means for management and

control. The power afforded by satellite observations would be beneficial if a government

possessed it for itself, but not if satellite data were acquired by an enemy with exploitative aims.

101 OTA, UNISPACE ‘82, 80. 102 “1982, UNISPACE Conference 1v,” PostBeeld, accessed 20 Apr 2020, https://www.postbeeld.com/sch1809- unispace-conference-1v. 103 Fan Zhang kindly translated the Chinese characters.

121

Figure 2.1. UNISPACE ‘82 stamp from the People’s Republic of China, 1982. (Reprinted with permission from PostBeeld)

Developing countries did not react to specific technical details of Global Habitability’s satellites because there were no technical specifics given. Indeed, that was the problem. In the absence of details, countries could not be sure how satellite data would be collected and accessed. In the absence of detailed specifications, there was a (not wholly unreasonable) suspicion that old colonial asymmetrical power relations would dictate data dissemination and use. The Global Habitability workshop report mentioned that unspecified satellite data would be collected from a fleet of satellites observing the Earth’s oceans, lands, atmosphere, and ice. It mentioned a few example satellite missions already in operation or in the planning stages (e.g.

Landsat, UARS, TOPEX). But in the development of its scant 13-page proposal, participants at the Global Habitability workshop completely failed to anticipate concerns about remote sensing data collection and dissemination. The workshop report contained not a single mention of the need for data regulation policies, apart from asserting the importance of creating usable global

122

data sets.104 In the time and space available, it was not possible to specify precisely what instruments the satellites would carry, the specific kinds of data that would be collected, or who would control the collected data. This aroused suspicion rather than support at UNISPACE ‘82 because the conference was expressly framed to promote the practical benefits of space science and technology for all countries.

Instead of engaging in an international consultation process, the US developed Global

Habitability based on what a group of US scientists and government officials desired from an

Earth observing satellite program. It was publicly announced without any prior discussion or

consultation at UNISPACE ‘82. Other countries were invited to participate by providing in-situ

measurements but that was all. The US also did not ask for permission to proceed with the

initiative, nor did it solicit opinions about its feasibility or its specific areas of focus. It might be

countered that there was no widespread consultation simply because, as a matter of

circumstance, there was a narrow time frame to put together the project proposal before

UNISPACE ‘82. Yet even if more time were available, it is not clear that the US would have engaged in broader consultations. A month before UNISPACE ‘82, Kenneth Pedersen, NASA’s

Director of International Affairs, recommended that Global Habitability not be presented as an

internationally-led program since that, “might serve to restrict and even dilute the science

objectives of the program.”105 Global Habitability was developed by intention unilaterally.

Further, since it was used as a way to direct UNISPACE ‘82 discussions away from the topic of

the militarization of space, other countries viewed Global Habitability as nothing more than a US

104 JPL, Global Change, 12. 105 Kenneth Pedersen to Associate Administrator of Space Science and Applications, 1 Jul 1982, Global Habitability File, Box 18045, NASA HRC, Washington, DC.

123

attempt, “to offer the Third World a few crumbs of comfort” and to deflect from what other

nations thought was a key conference issue.106

In damage recovery mode, the OTA’s post-UNISPACE ‘82 report completely reversed

its position from its pre-conference guidance. Before UNISPACE ‘82, the OTA had warmly

embraced the project as politically useful, confident that the project would gain international

support. Its post-conference report began by observing that US reluctance to engage in

multilateral consultation had stemmed in part from concern over what US representatives viewed

as the “increasingly politicized” nature of UN-sponsored conferences since the late 1960s. The

OTA explicitly stated that the US wished, “to avoid a confrontation on the basis of the New

International Economic Order.”107 The OTA’s post-hoc analysis admitted that, with the benefit

of hindsight, the US made a mistake in its failure to engage other countries in pre-conference negotiations and workshops regarding its initiatives: “It might have been politically desirable for the United States to seek joint sponsors for its proposals or at least to involve other countries in a debate on their merits during the course of the conference.” The post-conference OTA report acknowledged that, “the developing nations were wary of the possible negative impact of space technology on their nations [sic] interests even in communications and remote sensing….Concern was also expressed that the ‘open skies’ policy of the United States, i.e.

nondiscriminatory access to remote-sensing data, would disclose secret information about the sensed country to its adversaries...The United States must address these inherently political

106 Marsh, “Only the Rich,” 473. 107 OTA, UNISPACE ‘82, 32 [emphasis added].

124

issues.”108 In retrospect, the OTA acknowledged that it was particularly important for projects

like Global Habitability to engage in international collaboration to ensure their success.109

The backlash against Global Habitability at UNISPACE ‘82 materialized because the

project’s framers made assumptions about the beneficial and benign character of the project,

presuming that space activities and Earth science research were indisputable goods “for all

mankind.” The scientists framing Global Habitability thought of large-scale data collection and storage as an unequivocal good. Hence the pre-conference optimism, and then the post-

conference shock, forcefully captured by Shelby Tilford’s amazement at having been so

summarily “shot down.” This stark surprise resulted from a double blindness. The first

blindness was that groups like NASA’s Office of International Affairs and the OTA believed that

US unilateralism in international scientific projects was well warranted to avoid “undue”

politicization and to avoid “restriction” or “dilution” of scientific and technological activities.

The OTA was candid that the US was not interested in debating the militarization of space at

UNISPACE ‘82, nor did it wish to engage with concerns about economic security and national

sovereignty prompted by the recently articulated NIEO. The second blindness was that Global

Habitability’s science team, as well as NASA and government officials, all failed to account for

a critical political and economic security issue newly important to developing countries: data

ownership and control.

CONCLUSION

After 1982 NASA fortuitously benefited from changing international attitudes towards remote

sensing data. As Earth observing satellites became more common, countries tended to view their

108 Ibid, 73-4. 109 Ibid, 53.

125 advantages as outweighing harms, and therefore concerns over data collection and dissemination diminished in relative importance.110 Further, the US’s more adversarial position with respect to the UN that arguably reached a peak during the Reagan administration effectively marginalized most UN bodies, the main world forums in which developing countries expressed their concerns.

This antipathy to the UN is indicated by the US’s refusals to sign the UN Moon Treaty in 1979 and the UN Convention on the Law of the Sea in 1982,111 as well as its withdrawal from

UNESCO in 1984.112 The 1970s and early 1980s was a brief period when developing countries exercised a new political power on the world stage and incorporated scientific data into broader economic and social restructuring plans that they hoped would result in more equitable global economic and political relations. Had Global Habitability been introduced, say, ten years earlier or later—when data concerns and the power of developing countries were at a nadir rather than a zenith—the initiative might have received a different international reception. Instead, the international political and economic contexts of 1982, combined with a naive view of the neutrality of scientific data, proved to be insurmountable obstacles for Global Habitability.

110 Mack, Viewing the Earth, 188. 111 The main reason for US refusal to sign UNCLOS stemmed from its disagreement with sections of Part XI of the convention, which restricts control over mineral rights in areas outside sovereign control (known colloquially as the “high seas”) via the International Seabed Authority. Since this would curtain US activities in the region, US officials refused to sign the convention. Developing countries sought to bring “” areas like the high seas and outer space under international control to restrict exploitation by more developed and economically prosperous countries like the US. See: Krasner, Structural Conflict, 10, 227-66. 112 Among its many concerns with UNESCO, a US official explicitly mentioned the “New International Economic Order” and the “new world information order” as reasons for the withdrawal. Subsequent US Administrations have taken different stances towards UN bodies. For instance, in 2003 the US re-joined UNESCO, and then in 2017 announced it would withdraw again, effective 31 December 2018. See: Congressional Research Service, “UNESCO Membership: Issues for Congress,” accessed 30 Apr 2019, https://www.everycrsreport.com/files/20031120_RL30985_e03af874e629604f513ea11ba97563f1b96a5e17.pdf; Bernard Gwertzman, “US is Quitting UNESCO, Affirms Backing for UN,” New York Times, 30 Dec 1983, accessed 13 Mar 2020, https://www.nytimes.com/1983/12/30/world/us-is-quitting-unesco-affirms-backing-for- un.html; “United States Withdrawal from UNESCO,” International Legal Materials 23, No. 1 (Jan. 1984): 221; United States, Department of State, “The United States Withdraws From UNESCO,” accessed 30 Apr 2019, https://www.state.gov/r/pa/prs/ps/2017/10/274748.htm.

126

Global Habitability was “shot down” because project scientists, NASA personnel, and

government officials all implicitly and unilaterally endorsed a positivist view of data as neutral

and beneficial. They did not recognize that all data has a political dimension, no matter how

scientifically useful. As a result, they failed to recognize the importance of the international

context for a global scientific and technological program where unilateral actions by the US were

viewed as perpetuating old colonial power dynamics. The case of Global Habitability illustrates

that creating an “informational infrastructure” for a large-scale Earth science research program

requires the convergence of a number of components that might include scientific expertise,

technical capabilities, organizational interest, financial resources, and international support.113

When one or more of these components fails to coalesce, a proposed scientific research program fails to materialize. This is, of course, a recurring narrative in the history of science and technology, but as this case study contends, it should not be framed in terms of a distinction between science and politics. Global Habitability was among the earliest proposals for a large-

scale Earth science research project that used global technologies like Earth observing satellites

to turn the whole Earth into an object of scientific inquiry. As Lorraine Daston observes,

scientific objects take work to make because they have material and social dimensions that

makes them “elusive and hard-won.”114 Sometimes they are not “won” at all.

NASA officials and US scientists learned a number of important lessons from the

UNISPACE ‘82 experience. These included the need to develop a clearly articulated and

adequately detailed scientific rationale for a large-scale research program, and the need to

113 Drawing on Paul Edwards' work in Vast Machine (2010), Gemma Cirac-Claveras uses the case study of France's Laboratory for Studies and Research in Space Remote-Sensing (LERTS) to demonstrate how an “informational infrastructure” was developed for remote sensing data from Earth observing satellites in ecosystems ecology. See: Gemma Cirac-Claveras, “Satellites for What? Creating User Communities for Space-based Data in France: The Case from LERTS to CESBIO,” Technology and Culture 59, no. 2 (Apr. 2018): 203-25. 114 Lorraine Daston, “Introduction: The Coming into Being of Scientific Objects,” in Biographies of Scientific Objects, ed. Lorraine Daston (Chicago: University of Chicago Press, 2000), 2.

127 incorporate many external scientists, engineers, and even politicians into the development process. In 1983, NASA’s Advisory Council formed the ESSC to rebrand and rework Global

Habitability into an Earth science research program to study the Earth as an interconnected system.115 Chapters three and four detail these extensive efforts to develop a whole Earth research program that was scientifically and politically palatable, as well as being publicly accessible. While the new Earth system science (ESS) research program would encounter resistance from certain Earth science communities, it was more successful than Global

Habitability because at least part of the ESSC’s work gained widespread support and traction.

Through its long-term development and promotion of ESS, the ESSC’s phrase for describing an interconnected planet—the “Earth system”—would become the predominant way that Earth scientists refer to the Earth as a scientific object.

115 Funding for and mention of smaller scale “global habitability” studies (note the lower case usage) can be found up to 1984, but only as a very minor part of NASA’s overall budget and activities. At least as late as 1984, Shelby Tilford was still giving presentations that referred to “Global Habitability.” See: Shelby G. Tilford, “Global Habitability and Earth Remote Sensing,” Philosophical Transactions of the Royal Society of London. Series A, Mathematical and Physical Sciences 312, no. 1519 (Jul. 1984): 115–8.

Chapter 3 The Earth System Sciences Committee: Constructing a Research Program

INTRODUCTION

Global Habitability’s poor reception at UNISPACE ‘82 challenges the narrative that it is simply

obvious and necessary that humans would eventually study the entire planet as an interconnected

system. The Earth as a system was not a ready-made scientific object merely “out there” waiting

for the requisite technologies to facilitate its study. Once there were satellites able to take global

measurements and computers capable of analyzing and storing that data, it was not inevitable

that scientists would study the planet holistically as a system with interacting components. If all

of this were actually necessary, then the Global Habitability failure becomes largely inexplicable.

Global science and large-scale scientific research programs, like any kind of science, require work to establish. Scientific objects require construction and the generation of consensus. They are built up out of an innumerable and contingent combination of practices, instruments, theories, personalities, institutional needs, political concerns, and public interests. Not all construction projects succeed, as the Global Habitability case study demonstrates. Indeed, had Global

Habitability successfully generated support and become more widely known, it is possible that scientists might have adopted the phrase “global habitat” to refer to the Earth as a scientific object. What is certain is that scientists today predominantly use the “Earth system” phrase to do this work, not “global habitat” or another alternative. The “Earth system” is used in a widespread and uncritical manner now, but the history of how this came to be is linked to a group of scientists and administrators in the 1980s who worked to develop a large-scale Earth science research program to study the Earth as a system. Like Global Habitability, their goal was

128 129

to develop a research program that transformed the whole Earth into an object of scientific

inquiry. Their attempt was arguably more successful.

This chapter focuses on the work undertaken by this group, an ad hoc committee of

NASA’s Advisory Council (NAC) called the Earth System Sciences Committee (ESSC). From

1983 to 1988, the ESSC developed the inchoate ideas of Global Habitability into something that

could, it was hoped by those involved, generate widespread appeal and adoption by scientific

communities, politicians, and the public. They promoted what they called “Earth system

science” (ESS), a new research approach that would connect the previously separate Earth

science disciplines into a larger framework. Via the collection of global measurements with

satellites and complementary in-situ observations, data analysis, and the construction of quantitative models of the entire Earth system, the ESS research program would attempt to obtain a scientific understanding of the whole planet.1 It would describe how the Earth’s

components and their interactions functioned, how they evolved over time, and their probable

future evolutions. In particular, ESS would focus on the prediction of global changes that would

occur on timescales of decades to centuries.

This dissertation argues that the roots of the Earth system concept can be traced

specifically to the ESSC’s development and promotion of its ESS research program. Chapter

three examines the “development” piece of this story, while the “promotional” work will be

covered in chapter four. The ESS research program encountered a number of obstacles by

generating controversy and sometimes confusion among members of the various Earth science

disciplines. However, the committee’s broader conceptual framework—the planet as the “Earth

system” with interconnecting subcomponents—gained scientific traction and popularity. With

1 For the ESSC, “in-situ” meant all non-space-based measurements, including remote sensing data collected from instruments flown on airplanes. See: Earth System Sciences Committee (ESSC), Earth System Science: A Program For Global Change: A Closer View (Washington, DC: NASA, 1988), 107.

130

the ESSC’s work, the entire Earth became a scientific object in Lorraine Daston’s sense of the

phrase, designated by the “Earth system.” This conception of the planet eventually gained a

“heightened ontological status,” transforming the Earth from being studied as a system to the

Earth becoming a system.2 Today, the Earth is a system for Earth scientists, and this conception

forms one of the foundations upon which they conduct their research. Building on Paul

Edwards’ work that identifies global data collection by Earth observing satellites and global

computer modeling as crucial in the transformation of the entire planet into a “system,” chapters three and four emphasize the importance of an additional element: the social negotiations that concretized this transformation and facilitated the spread of the “Earth system” concept.3 These

social processes form an additional crucial component required for the emergence and continued

existence of a new scientific object, the Earth system.

Chapter three focuses on the formation of the ESSC and its development of a global

Earth science research program called ESS. In work by Erik Conway, Eric Goldstein, W. Henry

Lambright, John McElroy and Ray Williamson, and Richard Leshner and Thor Hogan, the ESSC

appears as only a relatively small part of their narratives. The focus has been on NASA’s longue

durée Earth science research activities or NASA’s development of the Earth Observing System

(EOS) satellites.4 However, the previously unexplored material available from ESSC chair

2 Lorraine Daston, “Introduction: The Coming into Being of Scientific Objects,” in Biographies of Scientific Objects, ed. Lorraine Daston (Chicago: University of Chicago Press, 2000), 10. 3 Paul N. Edwards, “Representing the Global Atmosphere: Computer Models, Data, and Knowledge About Climate Change,” in Changing the Atmosphere: Expert Knowledge and Environmental Governance, eds. Clark A. Miller, Paul N. Edwards (Cambridge, MA: MIT Press, 2001), 31-66; Paul N. Edwards, A Vast Machine: Computer Models, Climate Data, and the Politics of Global Warming (Cambridge, MA: MIT Press, 2010); Paul N. Edwards, “The World in a Machine: Origins and Impacts of Early Computerized Global Systems Models.” in Systems, Experts, and Computers: The Systems Approach in Management and Engineering, World War II and After, eds. Agatha C. Hughes, Thomas P. Hughes (Cambridge, MA: MIT Press, 2000), 221-53. 4 Erik M. Conway, Atmospheric Science at NASA: A History (Baltimore: Johns Hopkins University Press, 2008); Erik M. Conway, “Bringing NASA Back to Earth: A Search for Relevance during the Cold War,” in Science and Technology in the Global Cold War, eds. Naomi Oreskes and John Krige (Cambridge, MA: MIT Press, 2014), 251– 72; Edward S. Goldstein, “NASA’s Earth Science Program: The Bureaucratic Struggles of the Space Agency’s

131

Francis Bretherton’s personal files at NCAR offers a new view of the activities of ESSC

members and external collaborators. It enables a more detailed and nuanced portrait of the group

than was previously possible. These documents demonstrate the great efforts undertaken by

ESSC members to spread their ideas and achieve support. Analyzing these documents helps

explain why the ESSC’s Earth system concept has become so dominant in the Earth sciences

today.

Chapter three begins by providing a broad overview of the ESSC. It unpacks how the

committee was established, what it was charged to do, who were its appointed members, how it

conducted its work, and the major results of its deliberations. Secondly, the chapter examines

how the committee constructed its “Earth system science” (ESS) research program. This

involved the need to define the “Earth system” that this “science” would study. Lastly, the

chapter focuses on two controversies that emerged during the ESSC’s multi-year deliberations.

The first involved friction between Earth scientists working in disciplines that focused on

processes of widely divergent time scales. This played out largely as a disagreement between

geologists and geophysicists on one side and representatives from the rest of the disciplines on

the other (e.g. oceanography, atmospheric science, and ecology). These two camps offered

different ways to define the Earth system, what to include within the system and what to exclude.

Mission to Planet Earth” (PhD dissertation, George Washington University, 2007); W. Henry Lambright, “Administrative Entrepreneurship and Space Technology: The Ups and Downs of ‘Mission to Planet Earth,’” Public Administration Review 54, no. 2 (Mar./Apr. 1994): 97–104; W. Henry Lambright, “The Political Construction of Space Satellite Technology,” Science, Technology, and Human Values 19, no. 1 (Winter 1994): 47–69; Richard B. Leshner, “The Evolution of the NASA Earth Observing System: A Case Study in Policy and Project Formulation” (PhD dissertation, George Washington University, 2007); John H. McElroy and Ray A. Williamson, “The Evolution of Earth Science Research from Space: NASA’s Earth Observing System,” in Exploring the Unknown: Selected Documents in the History of the US Civil Space Program: Volume VI: Space and Earth Science, eds. John M. Logsdon, et al. (Washington, DC: NASA, 2004), 441-73; Richard B. Leshner and Thor Hogan, The View From Space: NASA's Evolving Struggle to Understand Our Home Planet (Lawrence, KS: University Press of Kansas, 2019).

132

However, neither side questioned the existence of the “Earth system” and the debate arguably helped reify the concept. The other controversial aspect of the ESSC’s work was its attempt to draft an “implementation” strategy for ESS. I argue that this was problematic for two reasons.

First, the strategy prioritized shorter timescales over longer ones. Second, it privileged space- based observations over in-situ measurements, and so those working in disciplines that relied on in-situ techniques felt marginalized. Given that this prioritization had funding implications, it is not surprising that this aspect of the committee’s work proved contentious. These controversies help explain why it was, ultimately, the broader and vaguer Earth system concept, and not the more narrowly defined Earth system science research program, that gained widespread acceptance (chapter five).

EARTH SYSTEM SCIENCE: A PROGRAM FOR GLOBAL CHANGE

Before looking at the ESSC, how it developed its ESS research program, and the resulting controversies, it is useful to first outline ESS itself in more detail. In its final report—Earth

System Science: A Program for Global Change: A Closer View (1988)—the ESSC presented a new “research approach” for the Earth sciences called “Earth system science” (ESS). ESS was envisioned as a global Earth science program that would work towards building a more comprehensive understanding of Earth processes and their interactions. These processes ranged over various timescales. There were short-lived processes that might last only seconds, days, weeks, or years (e.g. atmospheric turbulence, atmospheric convection, volcanic eruptions, earthquake cycles, seasonal vegetation cycles, global weather systems). Then there were medium-term processes that occurred on timescales of decades to centuries (e.g. soil erosion, climatic conditions, carbon dioxide variations, ocean circulation, nutrient cycles). Finally, there were long-term processes that occurred on timescales of thousands to millions of years (e.g.

133 glacial periods, soil development, speciation, atmospheric composition, plate tectonics, mantle convection, mountain formation).5

The overarching “goal” of ESS was: “To obtain a scientific understanding of the entire

Earth system on a global scale by describing how its component parts and their interactions have evolved, how they function, and how they may be expected to continue to evolve on all timescales.” The ESS would be global in scope—focusing on changes observable on a regional or planetary scale rather than smaller spatial scales—but, given the immensity of the task at hand, not all timescales could be given equal priority. Therefore, the ESSC set a more narrow

“challenge” for ESS that focused the “goal” on a timescale more relevant to human lifespans, of decades to centuries. ESS’s “challenge” was: “To develop the capability to predict those changes that will occur in the next decade to century, both naturally and in response to human activity.”6 Meeting the “challenge” of predicting changes on medium-term timescales of decades to centuries was of “great urgency” given the many environmental problems facing the planet.7

According to the ESSC, the Earth required study as a complex and dynamic body in perpetual change using a “systems” approach. In this approach, the Earth would be treated as a system comprised of interconnected subsystems (the atmosphere, hydrosphere, cryosphere, geosphere, biosphere), where energy and material flowed between these subsystems. These flows occurred via what the ESSC called Earth processes, associations of phenomena that could be described by physical, chemical, or biological laws. Inputs to one subsystem were the outputs to others, and they were linked in feedback loops that could, it was hoped, be measured

5 ESSC, Closer View, 27. 6 Ibid, preface [emphasis added]. 7 Ibid, 16.

134

quantitatively. A novel feature of ESS, compared to previous Earth science work, was its focus

on interacting processes that traversed the Earth’s subsystems. While disciplinary work was still

important, ESS emphasized the need for overcoming traditional Earth science boundaries to fully

understand interconnected planetary processes. Many disciplines ran up against the limits of this

disconnected research, with new advances only possible by interdisciplinary examination of

these interactions. For instance, the ESSC argued that to better understand processes in the

world’s oceans required interdisciplinary knowledge of the ocean-atmosphere interface, the

influence of polar ice, and ocean biota productivity, which linked ocean processes to those in the

atmosphere, cryosphere, and biosphere.8

Interdisciplinary work was key for ESS research. For the ESSC, ESS was not a new scientific discipline but a new “approach” to formulating Earth science research topics “that constitutes a synthesis and reorientation of existing research currents.”9 ESS would focus research on areas that had not received attention because it had not previously been possible.

With tools like satellites and computer models, the complex interactions between different parts of the Earth were now more amenable to research. The interdisciplinary study of the Earth as a system would be denoted by the “concept of an Earth system.”10 In its broadest strokes, this was

what the “Earth system” meant to the ESSC and this idea would receive widespread acceptance

in Earth science communities and beyond. It was a conception of the planet that gave a name to

the growing semantic void that had emerged in the Earth sciences with the rise of technologies

that promised to empirically support planetary research for the first time (chapter one).

However, the ESSC confronted resistance to how it chose to define the boundaries of the Earth

8 Ibid, 12. 9 Ibid, 139. 10 Ibid, 24.

135 system. While no scientists debated that the Earth was a system, there was much disagreement over how to define this Earth system (see below).

For the prioritized timescale of decades to centuries, the ESSC developed a conceptual model that represented the structure of the Earth system, its subsystems, and the major flows of matter and energy in the system (Figure 3.1). The conceptual model of the Earth system depicted it as a “wiring diagram,” with large purple rectangles representing the two predominant subsystems at the decades-to-centuries timescale: the physical climate system and the biogeochemical cycles system. The physical climate system was comprised of subsystem processes (grouped in orange and green boxes) relating to atmospheric physics/dynamics

(dynamics, cloudiness, radiation), ocean dynamics (sea ice, mixed layer, open ocean, marginal seas), and the terrestrial surface moisture/energy balance (energy, snow, plant transpiration/photosynthesis). The biogeochemical cycles system was comprised of processes relating to marine biogeochemistry (production, particle flux, decomposition/storage, open ocean, marginal seas), terrestrial ecosystems (vegetation, decomposition, insoluble and soluble processes, nutrient recycling, plant dynamics), and tropospheric chemistry (troposphere, cloud processes, urban boundary layer). Arrows between the green and orange boxes indicated the flows of matter and energy required to describe the interactions between subsystems, with the attached blue boxes indicating the specific material flowing (e.g. carbon dioxide, methane), the energy change occurring (e.g. wind stress, heat flux), or the process that linked two subsystems together (e.g. river runoff, albedo). The brown ovals represented processes external to the Earth system, but from which materials and energy could enter the Earth system, or to which energy and materials could flow out of the Earth system (e.g. the solar system, soil development). To take a simple example, within the physical climate system, the terrestrial surface moisture/energy balance subsystem was linked to the ocean dynamics subsystem via river runoff, which was in

136 turn linked to the atmospheric physics/dynamics subsystem via albedo. Closing the loop, the atmospheric physics/dynamics subsystem linked back to terrestrial surface moisture/energy balance subsystem via precipitation. One can follow many such flows on the diagram that connected these subsystems.

The ESSC developed this conceptual model for two reasons. First, it was intended to broadly indicate how the different components of the Earth system interconnected. Second, it was meant to represent how the different Earth science disciplines related to each other. The

ESSC viewed the subcomponents of this model as, “computer subroutines incorporating detailed knowledge of the relevant processes provided by the traditional Earth-science disciplines.”11

This may account for why the conceptual model resembles a flowchart for a computer program, which is an abstraction of the coding for the program (see below). Also, in the model each subsystem could be viewed as the domain of a particular discipline, with the diagram providing the visual framework to show the areas where interdisciplinary work was thought to be most needed. The conceptual model showed how the Earth could be understood as an interconnected system, how the distinct Earth science disciplines were ultimately interconnected from a global perspective, and gestured at what a computer model might ultimately produce. For the ESSC, the main point of the conceptual model was that it, “provides both a motivation and a mechanism for promoting interactions among specialists in different disciplines.”12

11 Ibid, 30. 12 Ibid, 31.

137

Figure 3.1. Earth system wiring diagram from the ESSC’s Closer View (1988), depicting Earth processes occurring on timescales of decades to centuries. (Design by InterNetwork, Inc | Payson R. Stevens; reprinted with permission from NASA)

138

But this conceptual model was also intended to form the basis for numerical models that

could be used to quantitatively simulate key components of the present and past Earth and, it was

hoped, eventually provide predictions about future conditions on the planet. For the ESSC, an

inherent component of the ESS approach is, “the desire for a quantitative understanding of the

Earth system and the many interactions among its components, rather than merely qualitative

descriptions.” This involved specifying the state of the Earth system with a set of variables, and

then applying appropriate physical, chemical, and biological “laws” to these variables to relate

the present state to the past and the future. For certain subsystems, for instance ocean dynamics,

the ESSC claimed that the set of required variables was already well known and it was simply a

matter of making the, “compromises necessary to restrict the description [of ocean dynamics] to

a manageable number of degrees of freedom” in the development of numerical models. For

other subsystems, for instance terrestrial ecology, scientists still, “lack an adequate understanding

of the basic chemical and biological laws governing assemblages of complex organisms and do

not yet know what set of state variables would be adequate to describe the major dependencies

within the ecosystem.”13

The ESSC recommended a comprehensive research program to address the overarching goal and challenge of ESS. First, global observation of all significant variables would be required. The ESSC listed 56 global variables that required long-term measurements, to be

obtained predominantly from Earth observing satellites but would also incorporate

complementary in-situ measurements. A large part of the report was dedicated to outlining the

space platforms that would be required for ESS research. This included near-term (1987-1995) objectives like the “extension and enhancement” of current and already planned NASA programs

13 Ibid, 30-1.

139 like Earth Radiation Budget Experiment (ERBE), UARS, and TOPEX. It also included longer- term objectives for the global observing program (1995 and beyond) that focused around

NASA’s Earth Observing System (EOS) space platforms that were intended to collect many global variables simultaneously. Second, scientists would conduct data analysis and interpretation of global datasets using current theoretical knowledge to discern patterns that indicated functioning Earth processes. In this step, scientists would identify the important state variables for particular subsystems. Third, scientists would build on the analysis stage by constructing mathematical and numerical models of Earth system processes. These would be used to compute past and present Earth system conditions. Lastly, modeling results would be compared with archived Earth system conditions to verify their ability to reproduce those states.

It was hoped that future models of parts or the entire Earth system would be able to make accurate projections of global trends into the future.

These four stages were cyclically linked together in ESS’s research approach, with global datasets providing the material for analysis and interpretation as well as the data with which to verify the efficacy of numerical models. Data analysis and numerical modeling could, in turn, provide new insights into the global variables for which to collect measurements. For all of this to be possible, it would be necessary to develop a global information system capable of facilitating data reduction, analysis, and numerical modeling, along with more mundane, yet equally crucial, capabilities for data storage and access.14 Such large-scale research would be expensive, so the ESSC recommended the coordination of US federal agencies to fund and organize ESS research, as well as to foster the international connections that would link ESS

14 Ibid, 31.

140

with established international programs, for instance those undertaken by the World Climate

Research Programme (WCRP) and the International Geosphere-Biosphere Programme (IGBP).

With the proposal for ESS, for the first time scientists had outlined an empirical research program with specific scientific objectives and an implementation strategy for studying the entire

Earth. The Earth could now, with this ESS program, be the object of scientific inquiry in a way it could not for Humboldt or Vernadsky who had the ambition but lacked the technical capacity to obtain and assemble global datasets (chapter one). For the ESSC and those introduced to ESS

(chapter four), ESS was a specific research program that relied heavily on space observations and computer modeling, it focused on Earth processes occurring on timescales of decades to centuries on the global (rather than local) scale, and it was closely linked to NASA. ESS meant something specific, and proved controversial since its selection of timescale prioritized, and therefore marginalized, certain Earth science disciplines over others. To understand ESS, resistance to ESS, and how the Earth system concept became more widely accepted than the ESS research program itself requires an examination of how the ESSC was formed and conducted its work, as well as how it came to define both the Earth system and its ESS research program.

THE EARTH SYSTEM SCIENCES COMMITTEE

NASA’s Advisory Culture

That NASA’s next attempt, after Global Habitability, to develop a global research program achieved more success was in part due to the broad and fortuitous changes in attitudes towards the collection of remote sensing data by the international community.15 It is likely that almost

any new large-scale Earth science research program reliant on remote sensing data would have

15 Pamela E. Mack, Viewing the Earth: The Social Construction of the Landsat Satellite System (Cambridge, MA: MIT Press, 1990), 188.

141

received a different reception than Global Habitability did at UNISPACE ‘82 (chapter two).

That said, NASA officials and scientists also learned a number of lessons from their Global

Habitability experience. Specifically, there was a recognition of the importance of ensuring that

the scientific rationale for a new research program was clearly articulated and adequately

detailed, something the Global Habitability report lacked. As much as possible, external

scientists, engineers, and even politicians needed to be involved in the development process.

And lastly, significant attention had to be given to promoting a new research program to external

communities (chapter four). To do all this, NASA’s officials returned to a more traditional

source for advice on program development: the agency’s Advisory Council.

NASA has a long history of employing internal committees to provide advice and

counsel.16 In 1977, under presidential directive,17 NASA consolidated its various programmatic

advisory committees into the NASA Advisory Council (NAC).18 NAC’s functions were, “to provide advice and counsel to NASA management on the plans for, the work in progress on, and the accomplishments of NASA’s aeronautics and space program.”19 NAC’s findings and

recommendations were reported directly to NASA’s Administrator. Its membership consisted of

a rotating group of mainly non-NASA individuals usually from scientific or engineering

communities. However, NASA still considered NAC and its standing committees as, “‘internal’

16 Joseph K. Alexander, Science Advice to NASA: Conflict, Consensus, Partnership, Leadership (Washington, DC: NASA, 2017). 17 Memorandum from President Jimmy Carter to the Heads of Executive Departments and Agencies, 24 May 1977, File 16707, NASA HRC, Washington, DC. 18 The six committees were: NAC Aeronautics Advisory Committee, NAC Historical Advisory Committee, NAC Life Sciences Advisory Committee, NAC Space and Terrestrial Applications Advisory Committee, NAC Space Science Advisory Committee, and NAC Space Systems and Technology Advisory Committee. The heads of these six committees were to report to NAC and to NASA’s chief scientist. See: Background paper on NASA Advisory Council and Committees, 13 February 1980, NASA Advisory Council (NAC) Series, Box 16710, NAC Correspondence File, NASA Historical Reference Collection, Washington, DC. This organization has changed over the course of NAC’s history, and in 1981 the Space Science Advisory Committee became the Space and Earth Science Advisory Committee (SESAC). See: Alexander, Science Advice to NASA, 55. 19 Background paper on NASA Advisory Council and Committees, 13 February 1980, Box 16710, NASA HRC, Washington, DC.

142

advisory bodies, in the sense that they were chartered by NASA, their members were chosen by

NASA, and they provide their advice and counsel directly to NASA management.”20 Though

NASA considered NAC’s advice internal, it is more accurate to view NAC as a quasi-internal

organization, where the needs and desires of NASA officials encountered, challenged,

contradicted, bolstered, or amplified those of the external NAC members, with neither group

being wholly subservient to the other. NAC provided an opportunity for NASA officials to align

the agency’s programs and priorities with those of the wider research community. It also

facilitated interactions between external scientists and engineers serving as NAC members. Not

a mere tool for either NASA or external members to unilaterally shape the directions of NASA’s

programs, NAC provided an opportunity for the mutual alignment of internal and external

research objectives.

Even after the international failure of Global Habitability, many NASA officials were

still keen on formulating a large-scale Earth science research program for the agency, and they

turned to NAC for assistance. NAC’s interest in this kind of global research traces back at least

to June 1980, when it recommended such a program during the New Directions Symposium that

was held to brainstorm new priorities given the impending end of research and development for

the agency’s major program of the 1970s, the Space Shuttle.21 NAC meeting minutes throughout the early 1980s repeatedly recommended a large-scale Earth science research program.22 Burt

Edelson, Associate Administrator of NASA’s Office of Space Science and Applications (OSSA),

agreed. In a memo from August 31, 1982—only ten days after the end of UNISPACE ‘82—

20 Background paper on NASA Advisory Council and Committees, 13 February 1980, Box 16710, NASA HRC, Washington, DC. 21John Naugle, “Report on NASA Advisory Council’s New Directions Symposium, Woods Hole, Massachusetts,” 9 Sep 1980, File 16709, NASA HRC, Washington, DC. 22 “Summary Minutes of the NASA Advisory Council NAC Informal Task Force for the Study of the Mission of NASA, December 1–2, 1982,” 15 Mar 1983, File 16711, NASA HRC, Washington, DC.

143

Edelson tasked Pitt Thome of the OSSA to lead a conceptual study of a satellite system, without

any restrictions in size, mass, or power, “for all civil remote sensing R&D purposes: science--

geology, hydrology, ecology, atmospheric chemistry, climatology and oceanography; and

potential applications--agricultural assessment, renewable resource monitoring, nonrenewable

resource exploration, ocean monitoring, mapping, meteorology, etc.”23 The result of Thome’s

design study was “System Z,” a large-scale system which relied on a fleet of satellites collecting simultaneous measurements that would later be developed as the “Earth Observing System”

(EOS).24

In addition to the System Z conceptual study, NASA’s interest in global Earth science

research was made even more concrete in 1983 with the formation of an ad hoc NAC committee

tasked with developing the scientific and implementation details for such a program. According

to Shelby Tilford, (Director of Earth Sciences, OSSA), the US delegation had returned from

UNISPACE ‘82, “with our tail tucked between our legs,” convinced that the Global Habitability

project was not feasible. However, NASA Administrator James Beggs and Edelson were not

convinced. Tilford continued: “So we all got together and agreed that we would set up this huge

group of scientists from every aspect of Earth Science, get them together, and put together a plan

for what this could and would do.”25 With encouragement from Tilford and other NASA

officials, in May 1983 Edelson officially requested that NAC provide, “advice and counsel on

the future role, responsibilities, and implementation strategies for the Earth Science and

23 Memorandum from B.I. Edelson (Associate Administrator for Space Science and Applications) to Special Assistant to the Associate Administrator for Space Science and Applications, 31 Aug 1982, System Z File, Box 18048, NASA HRC, Washington, DC. 24 Conway, Atmospheric Science at NASA, 225. 25 Shelby Tilford, interview by Rebecca Wright, Washington, DC, 23 Jun 2009, NASA Johnson Space Center, “Earth System Science at 20 Oral History Project,” accessed 5 Jun 2019, https://www.jsc.nasa.gov/history/oral_histories/NASA_HQ/ESS/TilfordSG/tilfordsg.htm.

144

Applications program.”26 To this end, NAC proposed and NASA officials subsequently approved establishing the ad hoc ESSC.27

Forming the Committee

NASA officials certainly had their own motives for establishing the ESSC. These were mainly the prospect of developing its EOS satellites (previously System Z). Chunglin Kwa argues— drawing on Pamela Mack’s seminal Landsat study—that NASA’s experience with Landsat shaped its later support of EOS. Landsat’s focus on applications like resource monitoring rather than basic science meant that it was difficult for NASA to justify retaining control since the agency’s mandate was mainly for research and development, not operations. Indeed, NASA did not retain control of Landsat, which was transferred to NOAA in 1979, and then subsequently transferred to the private company EOSAT from 1985 to 1995. This experience, according to

Kwa, led NASA officials to throw support behind EOS, which was to be a large-scale Earth observation platform focused on fundamental science.28 Previous research by Conway,

Goldstein, and Leshner also highlights this aspect of the ESSC mandate, suggesting that it was formed to provide a scientific rationale for NASA’s EOS.29

26 Summary of the Subcommittee on Earth Science, Earth System Sciences Committee for NASA, [26 May 1983], ESSC/Bretherton Committee File, Box 18042, NASA HRC, Washington, DC. 27 Report on NASA Advisory Council Principal Council Recommendations and Actions, 1 Aug 1983, File 16710, NASA HRC, Washington, DC. It’s not clear when, exactly, the phrase “Earth system science” was conceived, nor by whom. Erik Conway credits its coining to Moustafa Chahine of NASA’s Jet Propulsion Laboratory (JPL). Chahine would eventually become an ESSC member and chair the Tropospheric Sounding and Imaging Working Group. See: Conway, “Bringing NASA Back to Earth,” 261; ESSC, Closer View, 202. What is clear is that even in its earliest stages of considering how to provide “advice and counsel” to OSSA on an Earth sciences program in early 1983, NAC was already using variations on this phrase (e.g. Earth System Sciences, Earth-system studies, Earth-system sciences). See: Summary of the Subcommittee on Earth Science, Earth System Sciences Committee for NASA, [26 May 1983], ESSC/Bretherton Committee File, Box 18042, NASA HRC, Washington, DC. 28 Chunglin Kwa, “Local Ecologies and Global Science: Discourses and Strategies of the International Geosphere- Biosphere Programme,” Social Studies of Science 35, no. 6 (Dec. 2005): 929–30. 29 Conway, Atmospheric Science at NASA; Goldstein, “NASA’s Earth Science Program,”; Leshner, “The Evolution of the NASA Earth Observing System.”

145

However, the ESSC was mainly made up of members external to NASA. Members came with their own opinions about the current state of the Earth sciences, the potential of new technologies for future research, scientific research priorities, and how best to achieve those priorities. ESSC members were elite scientists and engineers well entrenched in other US science organizations, and as such they would not have been easily manipulated by the aims and desires of NASA officials. None of the previous scholarship makes this claim, but by focusing primarily on NASA’s development of the EOS satellites, the role played by external agents in

ESS has been marginalized, if not entirely overlooked. NASA’s desires for ESS confronted the interests and concerns of non-NASA ESSC members. Inside and outside NASA, all were agreed about developing a global Earth science research program, but there was less consensus about the specifics of achieving that goal. As will be seen, non-NASA ESSC members contested both the logic and the appropriateness of aligning ESS scientific objectives and implementation strategy too closely with NASA’s internal EOS ambitions. Members and agency representatives made cases for both sides of this debate.

By the early 1980s, there was increasing consensus among Earth science practitioners that the components and processes studied by separate disciplines—in the atmosphere, hydrosphere, cryosphere, biosphere, and geosphere—were dendritically connected and studying these interconnection would produce fruitful research (chapter one). Evidence of this broad consensus came from the sheer number of similar committees, workshops, research programs, and reports that emerged throughout the 1980s, all with much the same theme and agenda. John

McElroy and Ray Williamson note that this was a time when science advisory committees played important roles in, “setting the agenda for Earth science and in defending that agenda in

146

the political process.”30 These groups included NASA’s Global Habitability workshop, the

National Academy of Sciences’ (NAS) Committee on Earth Sciences (CES), the World Climate

Research Programme (WCRP), and the International Council of Scientific Union’s IGBP. ESSC chairman Francis Bretherton explicitly linked these groups to the ESSC’s work. For him, “This variety of labels illustrates the widespread concern among scientists and others over the global environment of humankind, and the complexity of the task of better understanding it.” Though

these groups started out with, “a different set of assumptions about just what are the most

significant aspects of the system and about just what actions are realistic in addressing

them…[there is] a broad degree of consensus about fundamentals” that included the need for

interdisciplinary Earth science research to better understand the planet’s interconnected

processes.31

These various committees provided opportunities for experts from a wide variety of

communities to meet, share ideas, develop new ones, and then return to their home institutions,

bringing with them new perspectives. The committees themselves were a kind of

communication technology, facilitating synthesis and consolidation of current and future

research programs, and giving practitioners a direct role in advising policy makers and funders.32

The committees also functioned as a mechanism for building consensus amongst the sometimes

disparate communities, a role not altogether lost on individual committee members. In one of

the many humorous documents that emerged from the work of the ESSC and other NASA

advisory committees, oceanographers Mark Abbott and Seelye Martin compiled the “Words of

30 McElroy and Williamson, “The Evolution of Earth Science Research from Space,” 442. 31 Statement by Francis P. Bretherton at the House Committee on Science and Technology Hearings on the International Geosphere Biosphere Program, 12 Sep 1984, Folder 80, Earth System Sciences Collection, NCAR, Boulder, CO. 32 Thanks to Sverker Sörlin for raising this idea of conceiving of committees as a “technology” to me at the March 2018 annual meeting of the American Society for Environmental History.

147

Wisdom” uttered by the OSSA’s Dixon M. Butler. Butler played an important role in the conceptual development of both System Z and EOS. His “words of wisdom” were folksy and euphemistic phrases about committee work and report writing that Abbott and Martin translated into “plain english.” The document is organized as a “typical” committee report with an executive summary, introduction, body, and conclusions/recommendations. In their executive summary, the authors drily note the important role played by scientific committees. Because recent Earth science programs using satellite remote sensing platforms “cost a bundle of money,”

Abbott and Martin stated that committees were formed to, “document the scientific rationale of a program for the decision makers[.]” Further, “This plethora of committees also develop scientific consensus because eventually all of the members of the community are on at least one committee.”33 This kind of joking suggests that there was a keen awareness of the role of committees in communication and consensus-building.

Despite their usefulness, committees could also be a source of frustration and tedium.

This sentiment was expressed by another oceanographer, James Baker, who was one of the more active ESSC members. Baker shared a short story with other ESSC members called “Speed

Trap” (1967) by science-fiction writer Frederik Pohl. The story revolves around a scientist continually frustrated by the number of committee meetings and conferences he is forced to attend in order to hear and share new ideas, leaving him with scant time to actually conduct his own research. Ultimately it turns out that “otherworldly” forces actively prevent humans from achieving their full potential by keeping them busy with these committee meetings and

33 Mark R. Abbott and Seelye Martin, “Earth Observing System: Stakes in the Ground: An EOS Dictionary,” Oct 1986, Folder 153, Earth System Sciences Collection, NCAR, Boulder, CO [emphasis added].

148

conferences.34 That Baker passed on this story suggests both his personal identification with it

and suggests that he felt that others ESSC members would also sympathize.

Bretherton Takes the Helm

The ESSC joined the tediously long list of committees formed in the post-World War II era by

the US government and other institutions like the National Academy of Sciences to provide

scientific and technical guidance, often for interdisciplinary and large-scale projects that would

require extensive government funding. By July 1983, NAC had officially agreed to form the

ESSC and by November Francis P. Bretherton had been selected as its chairman.35 Originally

from Britain and trained at the in applied mathematics, in the 1960s

Bretherton transitioned his research attention towards atmospheric science and physical

oceanography.36 In particular, Bretherton focused on fluid dynamics and the coupling of

atmospheric and ocean models.37 This aligned well with the interdisciplinary character

envisaged for the ESSC. Shelby Tilford cited Bretherton’s research as the primary motivation in

choosing him as chair.38

There were a number of other factors that made Bretherton a suitable choice for chair.

He had served as the President of UCAR from 1973 to 1980, which managed NCAR. Since the

1960s, NCAR has been a hub for interdisciplinary research, where scientists studied the

34 Fred Pohl, “Speed Trap,” [1967], Folder 150, Earth System Sciences Collection, NCAR, Boulder, CO. 35 Memorandum from Stan Ruttenberg to Ray Arnold, 16 Nov 1983, Folder 123, Earth System Sciences Collection, NCAR, Boulder, CO. 36 Bretherton’s dissertation was on “The Motion of an Incompressible Liquid at Low Reynolds Number.” See: Mathematics Genealogy Project, “Francis Patton Bretherton,” accessed 5 Jun 2019, https://www.genealogy.math.ndsu.nodak.edu/id.php?id=101678. 37 See: F.P. Bretherton and M. Karweit, “Mid-Ocean Mesoscale Modeling,” in Numerical Models of Ocean Circulation: Proceedings of a Symposium Held at Durham, New Hampshire, October 17-20, 1972 (Washington, DC: National Academy of Sciences, 1975), 237-49; Francis P. Bretherton, “Ocean Climate Modeling,” Progress in Oceanography 11, no. 2 (1982): 93-129; John Wainwright, “Earth-System Science,” in A Companion to Environmental Geography, eds. Noel Castree, et al. (Malden, MA: Wiley-Blackwell, 2009), 149. 38 Tilford interview, 23 June 2009.

149

atmosphere and its interconnections with processes on land, in the oceans, and with the Sun

(chapter one).39 In 1974, Bretherton became the Director of NCAR as well as UCAR. He

retained both of these positions until stepping down in 1980 to resume his own research, staying

on at NCAR as a senior scientist until 1987.40 Further evidence of Bretherton’s interdisciplinary

capacity comes from his chairmanship of NASA’s National Oceanic Satellite System (NOSS)

Science Working Group beginning in the spring of 1980. Though NOSS was ultimately

removed from NASA’s budget, this group submitted a report to NASA regarding, “the scientific

needs of the oceanographic community as a whole[.]”41 Reflecting years later, one former senior

NASA official recalled that, “Francis [Bretherton] had this vision of tying together much more

than the oceans, and there was a struggle to keep him focused on the oceans.”42 In short,

Bretherton was a “big picture” thinker with a demonstrated track record for leading

interdisciplinary research.

Bretherton was also a natural leader. His “big” personality included a booming voice that

could, reportedly, penetrate the interior walls of the Mesa Laboratory, NCAR’s home.43 ESSC member and solar-terrestrial physicist Leonard Fisk described Bretherton as, “an amazing chair” despite, or perhaps even because of, his voice. He joked that, “People measure sound volume in

Brethertons. He’s a very loud speaker. So at some point you say, ‘Francis, you’re up to two

39 James R. Fleming, Inventing Atmospheric Science: Bjerknes, Rossby, Wexler, and the Foundations of Modern Meteorology (Cambridge, MA: MIT Press, 2016), ch. 5. 40 In 1988, Bretherton announced his acceptance of the directorship of Space Sciences Engineering Center (SSEC) at the University of Wisconsin-Madison. See: “Bretherton Accepts Wisconsin Post,” Staff Notes: National Center for Atmospheric Research, 6 Oct 1988, Francis P. Bretherton File, NCAR, Boulder, CO. 41 NASA/NOSS Working Group, Needs, Opportunities and Strategies for a Long-Term Oceanic Sciences Satellite Program (Boulder, CO: National Center for Atmospheric Research, 1981), v. 42 Personal correspondence. 43 Legend has it that Bretherton’s writing on a chalkboard was so loud that people in the adjoining room could hear (personal correspondence with NCAR staff).

150

Brethertons.’”44 Bretherton was a known entity at NASA. In addition to his NOSS work,

Bretherton participated in a NASA workshop on a climate observing system that might serve as

NASA’s contribution to study the climate as mandated by the US National Climate Program Act

of 1978.45 He also took part in the early preparations for Global Habitability and was a member

of the NASA Advisory Council’s Space and Earth Science Advisory Committee (SESAC) from

1982 to 1985, helping to develop the initial outline for what would become the ESSC.46

Bretherton was well-placed, had an interdisciplinary disposition, and a history of serving on

NASA-affiliated scientific committees and working groups.

Both Tilford and Bretherton himself recalled an initial hesitancy to take on the chairman position for the ESSC.47 Tilford recounts that, “After thinking about it for a while, he

[Bretherton] agreed that it would be a difficult job, but he thought it would be a very worthwhile thing to do.”48 Tilford does not elaborate on Bretherton’s reluctance, but presumably Bretherton

recognized the immensity of the task and he might have been concerned that it would be difficult

to formulate this research program in the two year timeline established by the ESSC Terms of

Reference.49 Bretherton’s own words provide some insight into his hesitation. In an invited

44 Leonard Fisk, interviewed by Rebecca Wright, Ann Arbour, MI, 8 Sep 2010, NASA Johnson Space Center “Oral History Project,” accessed 12 Aug 2019, https://historycollection.jsc.nasa.gov/JSCHistoryPortal/history/oral_histories/NASA_HQ/Administrators/FiskLA/Fis kLA_9-8-10.pdf. 45 Conway, Atmospheric Science at NASA, 213. 46 Meeting Request for “Long-Term Changes which might affect the Habitability of the Earth” Initiative, 8 Apr 1982, Global Habitability File, Box 18045, NASA HRC, Washington, DC; Recommendations of the Subcommittee on Earth Science, “Earth System Sciences Committee for NASA,” 26 May 1983, ESSC/Bretherton Committee File, Box 18042, NASA HRC, Washington, DC; Curriculum Vitae, Francis P. Bretherton, Francis P. Bretherton File, NCAR, Boulder, CO. 47 Tilford interview, 23 June 2009; Wainwright, “Earth-System Science,” 149. 48 Tilford interview, 23 June 2009. 49 Though, as one ESSC member suggested, perhaps Bretherton himself was in part to blame for the long time it took to produce the ESSC reports. Berrien Moore III stated that Bretherton, “just talks forever, and so it was going to take forever to get this thing [an ESSC report] done.” See: Berrien Moore, interviewed by Rebecca Wright, Norman, OK, 4 Apr 2011, NASA Johnson Space Center, “Earth System Science at 20 Oral History Project,” accessed 10 Jun 2019,

151

contribution to a 1985 special issue of the Proceedings of the Institute of Electrical and

Electronics Engineers (IEEE) on remote sensing in the Earth sciences,50 Bretherton noted that,

“The very attempt to articulate an intellectually coherent structure across a field so broad is itself perilous, and the implicit claim to influence over the future advance of knowledge is bold, to say the least.”51 In a 1988 article for the National Science Foundation’s (NSF) Mosaic magazine,

Bretherton acknowledged that his first reaction to being asked to chair the ESSC was, “It can’t

be done” due to the divides between the Earth science disciplines. According to Bretherton,

“They just don’t talk to each other as communities.” However, “after an hour on the phone”

with a NASA official, Bretherton “reluctantly agreed” to head the committee.52 Bretherton was

attracted to the project not just because he valued the promotion of interdisciplinary Earth

science work, but also because he recognized that the Earth science community needed to get

more organized in order to compete for government funding with better coordinated scientific

communities like those of astronomers and planetary scientists.53

The ESSC Mandate and Membership

In 1983, with Bretherton installed as chair, NAC gave the ESSC three immediate tasks. First, the

committee was to review the science for studying the Earth as a system of interconnected

https://historycollection.jsc.nasa.gov/JSCHistoryPortal/history/oral_histories/NASA_HQ/ESS/MooreB/MooreB_4- 4-11.pdf. 50 The specific origins of this special issue are unclear, but Bretherton was invited to contribute by guest editor David A. Landgrebe, an ESSC member and electrical engineer at Purdue University. See: Letter from David A. Landgrebe to Francis Bretherton, 20 Jul 1984, Folder 105, Earth System Sciences Collection, NCAR, Boulder, CO; David A. Landgrebe, “Scanning the Issue: The Special Issue on Perceiving Earth’s Resources from Space,” Proceedings of the Institute of Electrical and Electronics Engineers 73, no. 6 (Jun. 1985): 947-9. 51 Francis P. Bretherton, “Earth System Science and Remote Sensing,” Proceedings of the Institute of Electrical and Electronics Engineers 73, no. 6 (Jun. 1985): 1119. 52 Edward Edelson, “Laying the Foundation,” Mosaic 19, no. 3/4 (Fall/Winter 1988): 7. 53 Goldstein, “NASA’s Earth Science Program,” 133.

152

biological, physical, and chemical components in the atmosphere, oceans, land, ice, and biota.

They were to build on the Global Habitability report as well as other Earth science committee

publications in the late 1970s and early 1980s.54 Second, the ESSC was to develop a strategy to

implement a global Earth science research program and establish research priorities. Lastly, the

committee was to define NASA’s role in this kind of research. The original ESSC Terms of

Reference gave two-years to produce a final report.55 As it would turn out, this was an all-too

ambitious timeline, as Bretherton had anticipated. There would be two interim publications in

1986 (the Preview brochure and the Overview report), with the final report not published until

January 1988 (the Closer View report), almost five years after the ESSC’s formation. The 15-

member committee was comprised of university, government, and industry scientists and

engineers representing a broad swath of Earth science disciplines: atmospheric physics and

chemistry, physical and biological oceanography, ecology, geophysics, geology, cryology, and

solar-terrestrial physics.56 This was a witty, erudite, well-written (and presumably well-spoken)

group of elite practitioners, many of whom had already served on a variety of other science and

engineering committees in the 1980s.57 A further 239 individual scientists and administrative

54 “Appendix C” of the ESSC’s Closer View report lists a “Selection of Recent Reports Relevant to Earth System Science.” These reports include ones prepared by the National Research Council on using space technologies in the Earth sciences and reports on the IGBP. See: ESSC, Closer View, 195. 55 NAC Informal Earth System Sciences Committee Terms of Reference, n.d., ESSC/Bretherton Committee File, Box 18042, NASA HRC, Washington, DC. 56 As John Wainwright notes, there was no hydrologist on the committee as none that were identified to serve on the committee were available. See: Wainwright, “Earth-System Science,” 149. However, ESSC members were aware of this lack and were in contact with a hydrologist who provided them with a report to help them incorporate hydrological concerns into the final report. See: Letter from Robert M. Ragan to Ray Arnold, 7 Aug 1985, Folder 84, Earth System Sciences Collection, NCAR, Boulder, CO. 57 Sébastien Dutreuil, in his recent dissertation on the Gaia Hypothesis, notes the many criss-crossings of science and technology committee membership, both in the US and in international organizations. See: Sébastien Dutreuil, “Gaïa: Hypothèse, Programme de Recherche Pour le Système Terre ou Philosophie de la Nature?” (PhD diss., University of Paris 1 Pantheon-Sorbonne, 2016).

153

support staff contributed to the preparation of the ESSC’s reports, either through participation in

one of the ESSC’s working groups or in the process of report review.58

Electronic Mail: Bringing the Background to the Fore

Though the vast majority who contributed to the ESSC reports were scientists, other individuals

with different backgrounds played key roles for the ESSC. Previous scholarship on NASA’s role

in Earth science research and its Earth Observing System—work by Conway, Goldstein,

Lambright, McElroy and Williamson, and Leshner and Hogan—all rely primarily on some

combination of committee reports, NASA archival material, and interviews with scientists and

engineers.59 Not surprisingly, these stories focus on scientists and engineers, which is of course

important, but it is only part of the story. Some of the most crucial contributors have been

overlooked, including a project manager from UCAR (Laura Lee McCauley), an external

consultant (Paul Blanchard), and a graphic artist (Payson Stevens). Chapter four examines in

detail the roles played by Blanchard and Stevens in the aesthetic and rhetorical aspects of the

ESSC reports, everything from word choice and document layout to image selection and media

support plans. McCauley’s invaluable role is ever-present, even if usually in the background.

By examining Bretherton’s personal files, the contributions of non-scientists to the relative success of the ESSC becomes clear.

In particular, electronic mail messages bring to the fore the background world of managers and support staff that kept the ESSC functioning for almost five years, from inception

58 ESSC, Closer View, 203. 59 Conway, Atmospheric Science at NASA; Conway, “Bringing NASA Back to Earth”; Goldstein, “NASA’s Earth Science Program”; Lambright, “Administrative Entrepreneurship and Space Technology”; Lambright, “The Political Construction of Space Satellite Technology”; Leshner, “The Evolution of the NASA Earth Observing System”; McElroy and Williamson, “The Evolution of Earth Science Research from Space”; Leshner and Hogan, The View From Space.

154

to final report. Email—or electronic mail as it was still called at the time—was not in

widespread use in the 1980s. To use it, one either had to have access to an expensive mainframe

computer, or one could use a personal computer and simply purchase electronic mail service

through a provider, making use of the provider’s access to a mainframe computer. Further,

electronic mail communication could only occur between those using the same service

provider.60 Telenet was one of the first internet providers to offer an electronic mail service, which the company called “telemail.”61 As advertisements from the 1980s show, there was at

that time no sustained and widespread market for electronic mail. Potential customers still

required convincing. Promotional material for the telemail provider Omnet claimed that, “Far

from being a threatening ogre, electronic mail is one of the easy, new ways in which we can use

the computer as a tool, not as an end in itself.”62 This new communication technology promised

to radically transform the ease and speed with which geographically dispersed individuals

interacted and shared information. Most members of the ESSC communicated by email, but not

everyone was convinced. For instance, one ESSC working group member, Frank Richter of the

Earth Systems Modeling Working Group, expressly chose not to be on telemail.63 Richter was,

however, a minority among the ESSC and its collaborators.

Groups within NASA were early adopters of email, and by the early 1980s had

subscribed to telemail providers.64 Not long after the ESSC formed in 1983, ESSC program

60 This was before protocols were developed to allow emails to be sent to any email address, regardless of provider. See: Abbate, Inventing the Internet, 106-10. 61 Janet Abbate, Inventing the Internet (Cambridge, MA: MIT Press, 1999), 80, 203. Since the ESSC makes use of Telenet’s telemail communication system (via the Omnet service provider), I will refer to these communications as “telemail messages” rather than “emails” or “electronic mail.” 62 Robert Heinmiller, “Use Electronic ‘Mail’: Save Time and Money, Get More Work Done,” Jun 1983, Folder 106, Earth System Sciences Collection, NCAR, Boulder, CO. 63 Telemail from L. McCauley to J. Dutton, 13 Dec 1984, Folder 89, Earth System Sciences Collection, NCAR, Boulder, CO. 64 Personal correspondence, Stanley Wilson, Dixon Butler.

155

support staff Laura Lee McCauley procured telemail services for the group.65 The critical

decision to follow the lead of other NASA groups allowed ESSC members from around the

country to communicate in near real time. This proved particularly valuable for report editing.

Biogeochemical cycle specialist and ESSC member Berrien Moore III noted that telemail

messaging was, “right at the heart of building a community.” When reflecting on how the ESSC

might have functioned without telemail, Moore was unequivocal in his assessment: “I don’t think

that any of this would have worked. It just would have been too episodic; you couldn’t have

gotten from one meeting to the next, there wouldn’t have been enough connectivity. The space

for compromise would have shrunk rather than expanded. It just wouldn’t have worked.”66

From a historiographical perspective, the telemail messages provide a window into the inner

workings of the ESSC. The final reports present a unified facade that is belied by the many

differences and disagreements expressed in the back-and-forth telemail messages. By exploring these email discussions, previously unknown figures working with the ESSC move into the foreground of the committee’s activities.

Though the majority of the archival research for chapters three and four comes from files preserved by Francis Bretherton, the majority of the correspondence is to or from Laura Lee

McCauley, the ESSC coordinator and manager from UCAR. McCauley was a central actor throughout the ESSC’s existence. McCauley took all of the meeting and conference call notes, sent out reminders and messages to ESSC and working group members via telemail, made travel and presentation arrangements, organized ESSC written material creation through review, printing, and distribution, coordinated public relations material and events, and contributed to the report review processes. While obviously much of the scientific content of the ESSC reports

65 Agreement for Electronic Mail, 1 Apr 1984, Folder 130, Earth System Sciences Collection, NCAR, Boulder, CO. 66 Moore interview, 4 Apr 2011.

156

came from the scientists and engineers, the role that McCauley and other support staff played

cannot be ignored. Bretherton usually receives most of the credit for the ESSC’s work. This is

largely deserved, as Bretherton spent years as chair, compiling the necessary scientific material,

synthesizing it into reports, and giving presentations all around the world on the ESSC’s

activities. But McCauley was the keystone of the ESSC, the heart that continually pumped to

make sure that all bodily parts of the committee received the requisite amount of blood for

survival.

While working for UCAR in 1984, McCauley and a few other staff were assigned to

provide technical and scientific support services for the ESSC in a subcontract agreement

between UCAR and Systematics General Corporation (SGC)—SGC itself under contract with

NASA to provide technical, administrative, and word-processing services.67 Though humble whenever she was thanked or received praise, McCauley occasionally joked about her critical role in ESSC activities. When discussing a promotional strategy for the Overview report

(chapter four), she asked ESSC members to think about who they knew in the media and elsewhere that might be able to help with spreading the word about ESS. But McCauley cautions: “JUST DON’T DO ANYTHING RASH OR FRANCIS [Bretherton] WILL HAVE

MY HEAD (and then where would we be!).”68 In a striking (and gendered) telemail exchange

between ESSC member and atmospheric physicist John Dutton and the Omnet telemail provider,

67 Systematics General Corporation/University Corporation for Atmospheric Research Agreement, 28 Apr 1984, Folder 123, Earth System Sciences Collection, NCAR, Boulder, CO; Memorandum from Stanley Ruttenberg to Systematics General Corporation, 6 Jan 1984, Folder 123, Earth System Sciences Collection, NCAR, Boulder, CO. 68 Telemail from L. McCauley to ESSC Executive Committee, 7 Jan 1986, Folder 55, Earth System Sciences Collection, NCAR, Boulder, CO. The source for this is a telemail message. Today’s convention with respect to the use of capital letters in emails is usually to interpret capital letters as “yelling,” but in the early 1980s capital letters were a way of emphasizing a point when it wasn’t possible to underline or italicize.

157

Dutton suggested that Omnet contact McCauley directly about its specific inquiry since she is,

“the Den Mother of ESSC.”69

McCauley and other ESSC support staff established electronic mail communication

services, wrangled editorial comments from tens to hundreds of collaborators, ensured that the

graphic designers for the committee reports had contracts and got paid, kept an overzealous and

micromanaging boss happy, read hundreds (probably thousands) of report pages, organized an

overhead projector for the meeting room, made sure that there were always coffee and donuts for

conference breaks. These activities may not be as well known or prestigious as scientific

research, but they are all necessary. This is all part of science in action and it is an inseparable

part of what post-WWII science committee work involves. When committee members come

from different institutions and scientific backgrounds, the work of organizing and managing the

group’s activities are essential for the smooth running of discussions and work. Managers like

McCauley are indispensable for a committee’s success.70 Without this work, large-scale science

after WWII would look quite different and arguably might not have even been possible at all.

CONSTRUCTING EARTH SYSTEM SCIENCE

Standing on the Shoulders of Giant Reports

Earlier reports from NASA and NAS in the first half of the 1980s promoted an interconnected

view of the Earth sciences and the need for a global observation and data analysis program.

These reports provided the crucial “starting point” for ESSC deliberations. The committee spent

69 Telemail from J. Dutton to OMNET Service, 3 Apr 1986, Folder 127, Earth System Sciences Collection, NCAR, Boulder, CO. 70 Thomas Hughes argues that managers are not only crucial for the success of post-WWII large-scale scientific and technological projects, but that management issues often present more challenges than those emerging from research and development. See: Thomas P. Hughes, Rescuing Prometheus: Four Monumental Projects That Changed the Modern World (New York: Vintage Books, 2000), 5.

158 much of its first year reviewing these documents and ensuring that its final recommendations and implementation strategy for an ESS research program aligned with this earlier material.71 Space policy expert Richard Leshner describes the ESSC’s final report, Closer View (1988), as, “an almost encyclopedic amalgamation of past reports on the subject of global change research[.]”72

In addition to the sparse Global Habitability report, the ESSC synthesized many subsequent reports into its ESS program. This included the OSSA’s System Z conceptual study from 1983 and the 1984 report from the EOS Science and Mission Requirements Working Group. This last report was an important resource for the ESSC since it reiterated the argument for the need of an integrated, global observation system (without mentioning the problematic phrase “Global

Habitability”). Citing the increase in atmospheric carbon dioxide, ozone depletion, acid rain, and

El Niño as just a few examples, the EOS Working Group noted that, “Over the last decade or so several problems in Earth science have emerged which require a multidisciplinary approach….The key to progress on these and other interdisciplinary issues in Earth science during the decade of the 1990s probably will be addressing those questions which concern the integrated functioning of the Earth as a system.”73 The EOS Working Group’s report outlined the space hardware that could support these scientific goals, which eventually centered around a proposed fleet of large satellite platforms in sun-synchronous orbit carrying multiple instruments

71 Memorandum from Ray Arnold to ESSC Members, 9 Mar 1984, Folder 129, Earth System Sciences Collection, NCAR, Boulder, CO; Systematics General Corporation/University Corporation for Atmospheric Research Agreement, 28 Apr 1984, Folder 123, Earth System Sciences Collection, NCAR, Boulder, CO; ESSC, Closer View, prologue, 195; ESSC Working Framework, Sep 1984, ESSC/Bretherton Committee File, Box 18042, NASA HRC, Washington, DC. 72 Leshner, “The Evolution of the NASA Earth Observing System,” 158. 73 NASA, Earth Observing System: Science Mission Requirements Working Group Report, Volume 1, Part 1 of 2 ([Washington, DC]: NASA, [1985]), v.

159 collecting simultaneous data on the Earth’s topography, temperature, winds, ice sheets, and atmospheric chemistry and dynamics.74

The most relevant non-NASA reports came from a committee formed by the NAS. The

NAS’ Committee on Earth Sciences (CES) produced two major reports in the 1980s to establish the goals and priorities for conducting Earth science research from space. To make it more manageable, the CES broke this work into two parts, each appearing as a separate report. The first report (1982) examined the solid earth, continental geology, and ocean dynamics, and was known as the Solomon Report after its chair, geophysicist .75 The second report

(1985) focused on the atmosphere and its interactions with the solid earth, oceans, and biota, and was known as the Prinn report after its chair, atmospheric scientist Ronald Prinn.76 Both reports detailed the Earth science objectives to be achieved over a ten-year period. Like the NASA reports mentioned above, the NAS reports promoted the study of the Earth as an interconnected system requiring extensive interdisciplinary cooperation and global observations collected from

Earth observing satellites. However, both the Solomon and Prinn reports stopped short of recommending specific space missions. The ESSC’s Closer View lists a number of other

74 Details for EOS were provided by the Science Steering Committee for the Earth Observing System in their report From Pattern to Process: The Strategy of the Earth Observing System (1984). EOS was originally conceived of in this report as consisting of three large polar-orbiting platforms that would be mounted with a number of remote sensing instruments (the number of platforms would be increased to four by the end of the 1980s). These platforms were to be launched in the mid-1990s and would collect data for at least ten years. The steering committee listed thirty different instruments that would be grouped together on the three platforms: 1. the Surface Imaging Sounder Package (SISP); 2. Sensing and Active Microwaves (SAM); and, 3. the Atmospheric Physics and Chemistry Monitors (APACM). The instruments would collect a wide-assortment of images of the Earth, including data in the visible, infrared, and microwave areas of the electromagnetic spectrum. See: Earth Observing System Steering Committee, NASA, From Pattern to Process: The Strategy of the Earth Observing System, EOS Science Steering Committee Report Vol. II ([Washington, DC]: NASA, [1984]), 10-12; Conway, Atmospheric Science at NASA, 226- 7. 75 Committee on Earth Sciences, Space Science Board, National Research Council, A Strategy for Earth Science from Space in the 1980s, Part I: Solid Earth and Oceans (Washington, DC: National Academy Press, 1982). 76 Committee on Earth Sciences, Space Science Board, National Research Council, A Strategy for Earth Science from Space in the 1980s, Part II: Atmosphere and Interactions With the Solid Earth, Oceans, and Biota (Washington, DC: National Academy Press, 1985).

160 relevant reports, but it was the Solomon and Prinn reports that served as the conceptual foundation for ESS.77

With so many committees and reports already in existence, it might be asked why it was necessary to form another committee to write yet another report to offer many of the same recommendations. Happily, ESSC member John Dutton provided a succinct answer.78

Responding to a request for comments on a draft of the ESSC’s Working Framework (see below), Dutton complained about its dry style and insubstantial content. He reminded

Bretherton of the need for the ESSC’s work to generate excitement and demonstrate why this research program would be, “worth half a billion a year of the taxpayers’ money[.]” Dutton continued: “I fear that this [ESSC Working Framework] is about as exciting as the

Solomon/Prinn reports, and it is in large part their blandness that is forcing us to do this job.”79

Though much of the basic scientific content of the ESSC’s reports had already been covered, the

ESSC’s contribution was to develop and present an integrated Earth science research program in a way that was scientifically rigorous yet compelling to non-specialist audiences. In particular, it was essential to generate Congressional excitement, and therefore approval and funding. They needed to “sell” their ideas in a way that previous committees and reports failed to do (chapter four). Further, the ESSC would not simply give a shopping list of Earth science objectives, but would also recommend an implementation strategy that described the kinds of missions that could achieve these objectives.

77 The Closer View lists 18, “recent reports relevant to Earth System Science,” published between 1980 and 1987. See: ESSC, Closer View, 195. 78 Dutton’s authorship is inferred from a signature. See: Letter from [John Dutton] to Francis [Bretherton], 26 Aug 1984, Folder 80, Earth System Sciences Collection, NCAR, Boulder, CO. 79 Letter from [John Dutton] to Francis [Bretherton], 26 Aug 1984, Folder 80, Earth System Sciences Collection, NCAR, Boulder, CO.

161

The ESSC’s Speed Trap: Committee Meetings and Report Delays

Originally, the ESSC planned to hold a series of meetings from December 1983 up to the end of

1985 to prepare its final report. It took quite a bit longer than initially expected. The ESSC’s

final report, the Closer View, was not published until January 1988. Over the course of the four- plus years of its existence, the ESSC met as a whole on ten occasions, three of which were multi- day summer studies that had become increasingly popular since the end of World War II.80

These meetings were attended by ESSC members, government agency and technical liaisons

(usually from NASA, the National Oceanic and Atmospheric Administration [NOAA], NSF, and

the US Geological Survey [USGS]), management staff (from NASA, SGC, and UCAR), and

invited observers (like Burt Edelson).81 There were also smaller meetings of the ESSC’s

executive committee, a group of the most highly involved members, as well as various writing

groups formed in the periods leading up to the publications of the Overview and Closer View.

When away from face-to-face meetings, the ESSC’s telemail service enabled members to share

everything from concerns and frustrations about specific issues to comments on draft reports and

opportunities to spread ESSC ideas to other individuals. All of this captures the raw amount of

work carried out by the ESSC, and suggests why members might have been receptive to Pohl’s

“Speed Trap.”

Early meetings and telemail report reviews culminated in the compilation of the ESSC’s

Working Framework in September 1984. The group’s first summer study was held in June 1984,

at Birdwood Pavilion, a former plantation house in Charlottesville, Virginia now used by the

University of Virginia as a conference center. Photographs of the meeting show cheerful,

80 Hughes, Rescuing Prometheus, 24. 81 Participant List as of 8 June 1984, Fourth Meeting, ESSC Summer Study, Charlottesville, Virginia, 8 Jun 1984, Folder 28, Earth System Sciences Collection, NCAR, Boulder, CO.

162

relaxed participants either sitting in meeting rooms with portable computers and overhead

projectors, or socializing over drinks in the evenings.82 Near the end of the summer study,

Bretherton reported to Edelson on the ESSC’s progress:

We [the ESSC] are firmly convinced that the time has come to study the earth as a system, and that space remote sensing will be central to that effort. The tools and the people are ready, the funds to do the job will not be astronomical, and, most important, critical scientific questions about the earth and its future can be posed and attacked….All in all, I am pleased to report to you that the Committee remains unanimously enthusiastic about the future of Earth System Sciences and about NASA’s role in that enterprise.83

There were two major outcomes of the Charlottesville summer study. First, attendees agreed on

the need to form working groups to focus on specific topics that required expert attention. These

groups would incorporate scientists from outside the ESSC and so might serve as tools for

consensus-building. Second, drafting groups met to “distill” the recommendations from the

reviewed reports. Bretherton used these summaries as he prepared the ESSC’s Working

Framework for September 1984.84 A summary of the Charlottesville discussions makes very

clear that this Working Framework was not an interim or draft report, “but primarily to be

regarded as a vehicle to present key concepts as seen at this time and to set the stage for

meaningful discussions.”85 As such, it was more a general roadmap for the ESSC activities,

which were to culminate in a final report by the end of 1985 or early 1986. The working

framework laid out in clear, if somewhat dry, language the five broad reasons why the ESSC’s

work was necessary: the need to increase focus on discipline-transcending questions in the Earth

sciences; the interaction between the land, waters, atmosphere, and biota at the global scale; the

82 Folder 145, Earth System Sciences Collection, NCAR, Boulder, CO. 83 Letter from Francis P. Bretherton to Burton I. Edelson, 14 Jun 1984, Folder 105, Earth System Sciences Collection, NCAR, Boulder, CO. 84 Draft Summary of Discussions, Earth System Sciences Committee (ESSC), NASA, Summer Study 11-15 June 1984, Charlottesville, Virginia, ESSC/Bretherton Committee File, Box 18042, NASA HRC, Washington, DC. 85 Ibid.

163

complexity of these dynamic interactions; the long-term and international requirements for

making global observations; and the timeliness of obtaining a better understanding of how the

planet functions on a global scale given the potential for human-caused environmental crises like

global warming and stratospheric ozone depletion.86

The Working Framework provided the guidelines that directed the ESSC’s work over the next couple of years, including another summer study in June 1985 at a resort on Orcas Island in

Washington State. With detailed working group reports in preparation, the ESSC’s primary efforts focused on drafting its final report. Though there were 15 ESSC members in total, a smaller core group—comprised of Bretherton, James Baker, John Dutton, and Berrien Moore, along with McCauley from UCAR and the consultant Paul Blanchard—led writing the final report. The first draft appeared in March 1985,87 with the writing group undertaking numerous

subsequent revisions in “frenetic activity all summer” in 1985.88 Group members wrote sections,

shared these sections with others (usually via telemail), and then Blanchard took these comments

and integrated them into a subsequent draft. This process continued into the fall of 1985.89

Among all the management staff that provided crucial assistance to the ESSC, Blanchard

is one of the more elusive actors. The results of his work are observable in the texts of the ESSC

reports and presentation slides, and he is copied on almost all important telemail correspondence.

But there is almost no surviving correspondence directly from him. Blanchard’s voice is almost

wholly absent. The origins of his involvement with the committee is something of a mystery, but

Blanchard played a central role in drafting and finalizing the ESSC reports. Founder and

86 ESSC Working Framework, Sep 1984, ESSC/Bretherton Committee File, Box 18042, NASA HRC, Washington, DC. 87 Outline for ESSC Final Report, [1985], Folder 4, Earth System Sciences Collection, NCAR, Boulder, CO. 88 Earth System Sciences Committee, Status Report on Major Issues, Francis P. Bretherton, [1985], Folder 73, Earth System Sciences Collection, NCAR, Boulder, CO. 89 Earth System Sciences Committee Calendar, 15 Jun 1985, Folder 18, Earth System Sciences Collection, NCAR, Boulder, CO.

164

president of the consulting firm Space Research and Management Inc., the Closer View report

lists Blanchard’s responsibilities as including, “Report Structure and Development, Text

Development and Editing, and Figure Development[.]”90 With a PhD in astronomy from

Harvard University and an MBA from MIT, Blanchard’s specialty was communicating scientific

and technical information to non-expert audiences.91 This made him particularly well suited to

help the ESSC achieve its objectives (chapter four). In addition to synthesizing comments into

the ESSC’s report, Blanchard also spent the summer and early fall of 1985 writing what was

planned as the Executive Summary for the final report.92

As late as July 22, 1985, the ESSC planned on having a final report draft ready to show

NASA by the end of September.93 However, by the fall of 1985, after numerous drafts and revisions and reviewer comments, the core writing group recognized that the original two year timeline laid out by NASA and NAC could not be met. A final report that included a detailed scientific rationale for ESS research and a concrete implementation strategy was not even close to being ready. Discussions at the ESSC’s seventh full meeting at NASA’s Goddard Space

Flight Center in Greenland, MD in September 1985 revealed that, while there was broad agreement about the contents of the final report, many of the finer details required significant further work. McCauley’s summary of the discussions records over 30 issues, ranging from minor corrections to major controversies. Bretherton reported to Edelson—a visitor at the meeting—that more time was required to properly “flesh out” ESS and to, “make sure that the drafts are circulated and given community review[.]” Edelson accepted this but requested an

90 ESSC, Closer View, 203. 91 “Paul Blanchard,” accessed 2 Jul 2019, https://www.crunchbase.com/person/paul-blanchard-4#section-overview. 92 Earth System Sciences Committee Calendar, 15 Jun 1985, Folder 18, Earth System Sciences Collection, NCAR, Boulder, CO. 93 Notes on “Preparation of Rept - Sched.,” [1985], Folder 8, Earth System Sciences Collection, NCAR, Boulder, CO.

165

interim report be published early in 1986 to present the ESSC’s major findings. What Blanchard

had been writing as an Executive Summary for the final ESSC report became an interim report,

the Overview, published in May 1986.94

The Ins and Outs of the Earth System

The ESSC’s first requirement for its Earth system science research program was to define Earth

system science. In a broad sense, this could be done easily enough. Bretherton offered an early

public sketch in his invited article for the IEEE. According to Bretherton, Earth system science

focused on building knowledge of the processes and interconnections of the Earth’s physical,

chemical, and biological components rather than studying these parts in isolation. Bretherton

stressed that many environmental processes—for instance the changes brought about by human

actions like the increase of carbon dioxide in the atmosphere from the burning of fossil fuels and

deforestation—needed, “an intellectual framework and a long-term program of research and

observations which transcend the traditional boundaries of the disciplines in Earth Sciences.”95

In its final report, Closer View, the ESSC stated that ESS is a “new approach” that studies the,

“Earth system...as a related set of interacting processes, rather than as a collection of individual

components. In anticipation of deeper insights into the interactions among these components,

Earth system science utilizes global observing techniques, together with conceptual and numerical modeling, to investigate both Earth evolution and global change.”96 ESS would study

94 ESSC Update No. 4, [1 Oct 1985], Folder 71, Earth System Sciences Collection, NCAR, Boulder, CO. 95 Bretherton, “Earth System Science and Remote Sensing,” 1119. Bretherton also uses the examples of the addition of chlorofluorocarbons (CFCs) to the atmosphere, the increase in acidity of precipitation, soil degradation, desertification, and water and air pollution from fertilizers. See: Bretherton, “Earth System Science and Remote Sensing,” 1118. 96 ESSC, Closer View, 13.

166

the interconnections between the Earth’s components rather than studying them in disciplinary

isolation, and it would do so using global datasets and numerical modeling.

In order to articulate a specific implementation strategy for ESS, the ESSC had to define

something new, the “Earth system.” What was part of the system and what was external to it?

Did it extend down to the center of the planet and up to the outer reaches of the Earth’s

atmosphere? Did it include the sun as well? Where should the Earth system’s boundary be

located? In a systems approach, the system under study needs to be defined. It is not given or

final or simply “out there” in an obvious or natural way. A system is not something inherent or

“there” in the world, but is just a label for whatever collection of things and relations is being

studied. In Rylean fashion, there is not another thing, “the system,” above and beyond the things

and relations.97 —coauthor of the (in)famous Limits to Growth (1972) that used systems analysis and modeling to predict ecological collapse if economic growth continued unchecked—defined a system as, “a set of things—people, cells, molecules, or whatever—

interconnected in such a way that they produce their own pattern of behavior over time.”98

Something about a system functions together, but its boundaries are defined by the

researcher, who sets the unit of analysis. Generally, what’s considered “internal” to the system

are just those components and processes that affect and are affected by other parts of the system,

and what is “external” are those things that can affect the system but are not affected by it.

System boundaries are not arbitrary labels placed just anywhere based on whim or fancy. They

are, however, conventions. System boundaries are placed at certain locations for reasons, though

they could be placed elsewhere for different but equally valid reasons. System boundary lines

97 Gilbert Ryle critiqued the Cartesian separation of “mind” and “body” by arguing that, contra the “dogma of the Ghost in the Machine,” to treat mental processes as separable from physical processes was to make a “category” mistake. See: Gilbert Ryle, The Concept of Mind, Sixtieth Anniversary Edition (New York: Routledge, 2009), ch. 1. 98 Donella Meadows, Thinking in Systems: A Primer, ed. Diana Wright (White River Junction, VT: Chelsea Green Publishing, 2008), 2.

167 can be drawn in different places for scientifically sound reasons. If, for instance, one is interested in studying global annual temperatures over the span of a few decades, then processes occurring on a similar timescale like ocean carbon dioxide uptake and vegetation cover would probably be internal to the system, while much slower processes like plate tectonics would be external (if considered at all). What is considered in and what is considered out of a system depends on the kinds of questions being asked, what is being studied, and importantly, the timescale of the study.

A main objective of ESS was to understand past and present states of the Earth system, and it was anticipated that this would lead to the ability to predict future states. Achieving this goal required the use of both conceptual and numerical models of the entire Earth system. The

ESSC tasked the Earth Systems Modeling Working Group (ESMWG) with studying the current state of Earth science modeling and describing a program for the integration and refinement of these models. The ESMWG was arguably the most important and influential of the ESSC working groups given that the prospects for Earth system science would be dim if extant models could not be integrated. Chaired by atmospheric physicist John Dutton, it included other prominent members of the ESSC like Bretherton and Berrien Moore.99 Numerical models of the entire Earth system were the ultimate target, whereby all significant global processes and variables would be entered into a computer program as equations that represented various components of the Earth system and their interactions. By the 1980s, these kinds of numerical models were being used successfully in NWP and climate modeling through GCMs for the atmosphere.100 But modeling the entire Earth system was still a distant goal that might not even

99 Other potential members and consultants included Paul Crutzen and James Hansen, though neither ultimately agreed to join. See: Telemail from J. Dutton to L. McCauley, 2 Sep 1984, Folder 89, Earth System Sciences Collection, NCAR, Boulder, CO. 100 See: Edwards, A Vast Machine, ch. 6; Edwards, “The World in a Machine,” 221–53.

168

be attainable. Assessing whether it was plausible to integrate, or “couple,” existing modeling

techniques was a painstaking and time-consuming technical task. Some components of the Earth system, like atmospheric and oceanic circulations, were being modeled in isolation as well as in

coupled models. Other individual models (like that of the carbon cycle) were also available.

Modeling of other components (for instance the evolution of the distribution of forest species as

a function of climate, upper atmospheric chemistry, ocean ecology) were still being developed.

Even longer term processes like crustal motion and plate tectonics appeared, for the ESSC,

“susceptible to modeling.” However, the coupling of the most advanced models was still in its

early stages, and many feedbacks and interactions between subsystems were not clearly

understood.101 According to Bretherton, “we don't have a concrete idea of when such a global

model can actually be implemented. There is an active effort to glue together some of the parts

of the system, such as the physical atmospheric circulation and the physical ocean circulation,

but the further you get into the biological part of the system, the fuzzier everything gets.”102 The

prospect of combining all available models, along with future models for other components of

the Earth system, was at best only a distant possibility and at worst an impossibility. To reduce

“fuzziness,” the more immediate task of the ESMWG was to develop a conceptual model of the

Earth system, one that depicted the system both visually and descriptively. This conceptual work

would ultimately decide what was “in” and what was “out” of the Earth system.

101 Earth System Sciences Committee (ESSC) Working Group on Earth System Modeling, First Meeting, Washington, DC, Oct 1984, ESSC/Bretherton Committee File, Box 18042, NASA Historical Reference Collection (HRC), Washington, DC. 102 Arthur Fisher, “One Model to Fit All,” Mosaic, 19, no. 3/4 (1988): 59.

169

Planet Dynos

The extent to which ESSC members were self-aware about the conventional and negotiated character of deciding what would be included in the concept-map of the Earth system is indicated by a joke circulated among members of ESMWG.103 On October 11, 1984

paleoclimatologist John Imbrie of Brown University sent a memo to John Dutton.104 Imbrie was

a fellow member of the Earth Systems Modeling group with an irrepressible sense of humour.

His memo to Dutton is brief, enclosing a “news clipping” that Imbrie hoped Dutton and the other

members of the working group could help him “puzzle out.”105 The clipping had the attention-

grabbing headline: “Contact Made With Life in Outer Space.” A sub-headline declared: “John

Dutton, Chairman of a NASA Working Group, Given Credit for the Discovery of Humanoids with Short Lifespans and Fast Metabolism Living on the Earth-Like Planet Dynos.” The news clipping continued for eight pages, as the author detailed the conditions on Planet Dynos, some of the characteristics of the dynapeople living there, and some of their science. Dynos is an

Earth-like planet that, “has exactly the same dimensions and orbital configuration as the Earth, and revolves around a Sun of the same size and age as ours.” However, the dynapeople are,

“quite a bit smaller than humans” with a metabolic rate that is, “very much faster than our own.”

As a result, “the maximum lifespan of a dynaperson was about 40 hours,” comparable to the

103 Though quite a prominent part of the sciences, there is a dearth of humanities scholarship that focuses specifically on scientific humour. Not only is scientific humour not rare, but sociologists Michael Mulkay and G. Nigel Gilbert argue that joking and humour can be used as a way to gain insight into a particular scientific culture and serious aspects of science. Studying scientific humour can help to show how particular scientific cultures are constituted and what kinds of beliefs, practices, and problems its practitioners share. See: Michael Mulkay and G. Nigel Gilbert, “Joking Apart: Some Recommendations Concerning the Analysis of Scientific Culture,” Social Studies of Science 12, no. 4 (1982): 585-613. 104 Imbrie was well known for his contributions to a study that used ice core samples to empirically support Milutin Milanković’s theory that historical variations in climate were linked to orbital variations of the Earth. See: D.L. Evans, et al., “Variations in the Earth's Orbit: Pacemaker of the Ice Ages?” Science 194, no. 4270 (10 Dec. 1976): 1121-1132. 105 Memorandum from John Imbrie to John A. Dutton, 11 October 1984, Earth Observing System Series, Box 18042, ESSC/Bretherton Committee File, NASA Historical Reference Collection, Washington, DC.

170

“average housefly,” which meant that their, “natural cycles of hunger, menstruation, telephone billing and so on go forward about 22,000 times faster than our own.”106 Putting aside the not- so-subtle sexism—a reminder that the ESSC and its collaborators were almost all men—Imbrie described Dynos as a planet with conditions that were the same as those on Planet Earth for all practical purposes. But Dynos was inhabited by humanoids with quite different lifespans than humans on Earth.

The important feature of the joke is that it shows that at least ESMWG members were notably aware that their decisions about what was “in” and what was “out” of the Earth system constituted what the system would ultimately be. There was a fundamental recognition that the contingent length of an average human lifespan shaped how the Earth system would be characterized. There was no scientifically “objective” system to be unearthed. How the Earth system would be characterized was inevitably species specific. A different species, with a different lifespan and therefore different concerns, would define the system differently. The planet Dynos is Earth-like in all important aspects (it’s the same size, it orbits around a similarly- sized sun with similar luminosity). One might presume that the planetary science on Planet

Dynos conducted by the dynapeople would closely resemble the work done by human scientists on Earth. But as it turns out, the 40-hour life expectancy of the dynapeople had led to some rather different concerns, and therefore different scientific emphases. Dynos system science was not the same as Earth system science.

For instance, the dynapeople’s meteorology was much simpler than that on Earth. As a result of a different lifespan, and therefore different temporal perspectives and concerns, meteorology on Dynos and Earth had quite different priorities. It was, apparently, really just

106 Ibid.

171

geography: “A typical rainstorm [on Dynos] lasts a good fraction of a typical dynaperson’s

lifetime, and doesn’t move about all that much. This means that you can simply pick the climate

you like, and build a house there. Rather convenient!”107 Instead of measuring the current state

of conditions in the atmosphere and then determining how these conditions will change over time

using calculations representing the motion of fluids and other mathematized processes, the short-

lived dynapeople could simply focus their meteorological efforts on observing current conditions

for a given location (its temperature, pressure, humidity, precipitation, wind speed, etcetera). In

the 40 hour life of a dynaperson, less than two Earth days, conditions simply do not change all

that much, especially when compared to the roughly 657,000 hours the average human could

hope to experience in a lifetime.108 On Dynos, it was not nearly so important to identify and

mathematize the dynamic processes that result in future changes to the current state. The

dynapeople were primarily concerned with the 40-hour timescale.

Atmospheric physics on the two planets also had different emphases. Given their short

lifespans, the dynapeople were particularly interested in studying cloud physics. It was not just

that cloud physics on Dynos was the same as that on Earth, just better worked out. The cloud

physics on Dynos was actually different. In the fictional article, Bretherton “congratulated” the

dynapeople on their cloud physics. He asked how they had handled, “the problem of

turbulence,” a problem that scientists on Earth struggled with. The dynaperson reply was tinged

with condescension: “Oh, we gave that primitive notion up long ago...along with the equally

primitive notion of eddy diffusion coefficients.” These issues matter to Earth scientists because

107 Memorandum from John Imbrie to John A. Dutton, 11 October 1984, Earth Observing System Series, Box 18042, ESSC/Bretherton Committee File, NASA Historical Reference Collection, Washington, DC. 108 The average human lifespan for Americans born in 1984 was 74.7 years. See: “Americans born in 1984 had an average life expectancy…” UPI Archives, 30 Mar 1987, accessed 6 Sep 2019, https://www.upi.com/Archives/1987/03/30/Americans-born-in-1984-had-an-average-life- expectancy/1036544078800/.

172

they play large roles in determining the dynamics of the atmosphere and oceans. But since

dynamics (i.e. change over time) plays only a small role in the dynapeople’s understanding of the

atmosphere, turbulence and eddy diffusion coefficients distract from more important research

objectives and can be discarded as “primitive” notions. An annoyed fictional Bretherton

described the dynaperson response as, “just a bit arrogant and -- how shall I say it? -- rather

uninformed about how physics operates in the real world. Or at least the world as we model it at

our NCAR.”109 Bretherton, in the article, is offended by the idea that certain concepts of central

importance to Earth system modelers could be completely disregarded by short-lived humanoids.

Part of the joke here comes from Bretherton’s huffy assumption that “how physics operates in the real world” is a stable and constant thing throughout the universe. But “article” Bretherton quickly qualified what he meant by physics operating in the “real world.” What he meant was actually how the world is modeled at “our NCAR.” The dynapeople were, according to fictional

Bretherton, uninformed not about how physics operated in the “real world” but about how the world was modeled at NCAR on Earth.

The joke revolves around the understanding that there is more than one way to model the

“real world.” How the world is modeled depends on the point of view and concerns of those doing the modeling. Atmospheric turbulence is part of the Earth system but not part of the

Dynos system. This is not because of any physical difference on the two planets, but because the dynapeople lived short and were, therefore, not concerned with atmospheric dynamics.

Imbrie as author, and the ESMWG members who read the piece (and were presumably amused by it) understood their models to be deeply situated and shaped by at least an anthropocentric point of view. For the thing being modeled, whether it be the Earth system or the Dynos system,

109 Memorandum from John Imbrie to John A. Dutton, 11 October 1984, Earth Observing System Series, Box 18042, ESSC/Bretherton Committee File, NASA Historical Reference Collection, Washington, DC.

173

what is in and what is out will depend on what the practitioners care about, what they choose to

include in the system, what they chose to identify as their object of analysis. Imbrie’s joke relies

on the recognition of this for its humour, and thus reveals some of the background assumptions

about modeling that the ESMWG members held that would not otherwise be accessible.

Wiring the World

ESMWG members recognized that the Earth system was not a natural entity “out there” in the

world ready-made, but that it required construction and would be based on the specific interests

of Earth scientists. John Dutton, in preparation for the ESMWG’s first meeting in October 1984,

described their strategy for developing models of the Earth system as a “top-down approach” by

which they would determine the overall structure of the system, its boundary conditions, its

subsystems, the variables within the subsystems, how they interacted, and what kinds of

observations would be required to develop this large-scale model.110 The first step in this top-

down theorizing would be developing a conceptual model of the Earth system that detailed its

components and interactions. This model would be, for all practical purposes, the Earth system.

For a brief period, the ESMWG’s Earth system model was to incorporate processes on all

time scales. However, members recognized the difficulties that this would entail and so began

distinguishing between shorter term and longer term processes. At their second meeting in

February 1985 in Jackson, Wyoming, the ESMWG decided to focus on shorter term processes,

occurring on timescales of decades to centuries.111 These were processes and changes like:

climatic conditions, ocean circulation, sea-ice dynamics, the effects of vegetation on land-surface

110 System Modeling Working Group Preliminary Description of Task, 14 Jun 1984, Folder 89, Earth System Sciences Collection, NCAR, Boulder, CO. 111 Telemail from J. Dutton to F. Bretherton, et al., 27 Feb 1985, Folder 91, Earth System Sciences Collection, NCAR, Boulder, CO.

174 , ozone chemistry in the stratosphere, and the biogeochemical properties of vegetative canopies. Such a decision was the result of three major drivers. First, the ESSC recognized that human activities were dramatically altering planetary conditions. To better understand the effects of human activities like fossil fuel burning, land use change, and the emission of CFCs, the ESSC prioritized global changes and Earth processes occurring on timescales proportional to a human lifespan, timescales of decades to centuries.112 Second, ESMWG members for the most part came from disciplines focused on shorter term processes like atmospheric physics and chemistry, physical and biological oceanography, and ecology. Personal research interests to some degree influenced the choice of timescales. Lastly, one goal of the ESSC was the development of a single numerical model of the Earth system that could accurately represent past and current changes and make predictions about its future states. With this modeling goal, it became necessary to make practical choices about what could be included in the system and what had to be excluded because it was impractical to incorporate into a single model. After many discussions, the ESMWG determined that it was simply not possible—given the current state of knowledge, available data, and computational power—to include all of the Earth’s major global processes on all timescales in a single wiring diagram. Instead, the ESMWG chose to emphasize the only “urgent” timescale relevant to human lifespans and policy making.113

This choice of a decades to centuries timescale in turn dictated what Bretherton called

“the choice of objectives,” what would be incorporated into the Earth system. According to

Bretherton, “The emphasis on global feedbacks effective for change over decades to centuries provides criteria for excluding from the core model many phenomena of local or regional

112 This was also the timescale favoured by the Global Habitability initiative. See: JPL, Global Change. 113 ESSC, Closer View, 15.

175 significance and global processes which are too slow to cause significant changes.” 114

Bretherton continues: “basic geology and geophysics are not directly relevant on these time scales, except as aids in interpreting drainage and soil patterns.”115 Topography, formed by the incremental movement of the Earth’s lithospheric plates, would affect the drainage patterns for precipitation on timescales up to decades and centuries, but topography in this timescale was not an interactive part of the Earth system since it would not be perceptibly altered by precipitation.

Though geologists and geophysicists studied phenomena that are connected to or part of the

Earth, the processes they studied were, according to Bretherton, only useful to the extent that they could provide the boundary conditions in which the shorter term processes occurred or which affected those shorter term processes. They were not themselves part of the Earth system.

By making this choice, the ESSC was also choosing how to define the boundaries of the Earth system. As the Dynos joke suggests, the ESSC explicitly recognized the pliability of boundary placement for the Earth system and implicitly recognized that the Earth system was being construed with respect to anthropic interests and effects. It was the researchers who established the boundaries of the system under study, not any inherent or objective property of the natural world.

Having defined the Earth system on the timescales of decades to centuries, the ESMWG could begin development of a conceptual model of this Earth system, what they referred to as an

Earth system “wiring diagram” (Figure 3.1). This model would depict the structure of the Earth system, its components, and the Earth processes that linked the components together in feedback loops. The development of the diagram was a group effort. There was no projector screen in their meeting room, so the group projected overhead transparencies (what they called “view

114 Bretherton, “Earth System Science and Remote Sensing,” 1119. 115 Ibid, 1122.

176 graphs” or “vu graphs”) containing an image of the wiring diagram directly onto the wall.

Bretherton said that participants, “kept running up to make changes directly onto the wall, which got pretty messy. In fact we had to pay for repainting it.”116 They believed that the diagram would provide a useful starting point for a numerical model of the Earth system. Each box of the diagram represented a computer subroutine comprised of global state variables that are interconnected to other variables and other subroutines via Earth processes, or algorithms. The wiring diagram was a flowchart that mapped relations between “subroutines” in the system that together made the Earth “function.” Earlier representations of an interconnected planet like that offered by the US Committee for GARP (chapter one) still included features that might be seen on Earth (for instance clouds, waves, glaciers, and mountains). The ESSC wiring diagram included no such features. It represented the Earth abstractly as a kind of computer program.

While members acknowledged that the diagram was an early, simple, and incomplete representation of the major Earth system components and processes for decades to centuries, it could nevertheless provide a useful starting point for subsequent elaborations.117 In this sense, it was a scientific research tool, not a complete picture or map. However, the diagram was also designed with an organizational project management purpose. It provided a visual representation of how the different Earth science disciplines fit together into the overarching ESS framework.

It was intended to serve as a tool for achieving interdisciplinary cooperation, depicting how the components of the Earth fit together. The ESMWG ultimately developed two versions of this

Earth system model, one simplified (Figure 4.5) and one more complex (Figure 3.1). The

116 Fisher, “One Model,” 59. 117 ESSC, Closer View, ch. 3 and 6.

177

simplified diagram—now known as the “Bretherton diagram”—became one of the most

influential and well-known products of the ESSC’s work (chapter four).118

Though the ESSC (and ESMWG) call their diagrams “wiring diagrams,” the

representations have similarities not just with circuit wiring diagrams in use since the nineteenth

century but also with flowcharts or flow diagrams that are used to represent structural

relationships, algorithmic steps, and/or processes. The earliest flow charts—called “process

charts”—were developed by Frank and Lillian Gilbreth in the early 1920s to represent (and

hopefully optimize) workplace processes.119 By 1949, John von Neumann and Herman

Goldstine had proposed the use of “flow diagrams” to depict operating instructions for

programming computers.120 These were also called wiring diagrams. According to historian

Mark Priestley, drawing computer programs as wiring diagrams was a common practice

beginning with the programming of ENIAC in the 1940s. Since the ENIAC’s switches and wires

required physical configuration before executing operations, its programs were written out as

electrical wiring diagrams rather than as lists of commands.121 This historical legacy of drawing

computer programs as flowcharts resembling wiring diagrams continued into the 1980s, at least

as a pedagogical tool.122 Given that the ESSC and ESMWG framed the long-term goal of ESS as being the development of numerical models that would be run on computers, it is unsurprising

118 In a number of the “ESS at 20” NASA oral histories, interviewees describe the Bretherton diagram as one of the most important, if not the most important, and lasting ESSC product. See: Dixon M. Butler, interviewed by Rebecca Wright, Washington, DC, 3 Jun 2010, NASA Johnson Space Center “Earth System Science at 20 Oral History Project,” accessed 10 Oct 2019, https://historycollection.jsc.nasa.gov/JSCHistoryPortal/history/oral_histories/NASA_HQ/ESS/ButlerDM/ButlerDM _6-3-10.pdf; Fisk, interviewed 8 Sep 2010; Moore, interviewed 4 Apr 2011. See also: Tim Lenton, Earth System Science: A Very Short Introduction (New York: Oxford University Press, 2016), 14. 119 Frank B. Gilbreth and L.M. Gilbreth, Process Charts (New York: The American Society of Mechanical Engineers, 1921). 120 Douglas Hartree, Calculating Instruments and Machine (Urbana, IL: University of Illinois Press, 1949), 112. 121 Mark Priestley, A Science of Operations: Machines, Logic and the Invention of Programming (London: Springer, 2011), 111. 122 Thanks to Greg Good who informed me of this pedagogical aspect.

178

that their Earth system conceptual model would take the form of a computer program wiring

diagram.

Another important historical source for the ESMWG diagram is the visual depiction of flows of matter and energy that emerged in the various Earth sciences beginning in the 1940s.

These representations originate with ecologist G. Evelyn Hutchinson’s visualizations of

biogeochemical cycles. A regular attendee of the Macy Conferences held from 1946 to 1953—

the conferences responsible for developing and promoting cybernetics—Hutchinson drew

diagrams to show how elements like carbon cycled through various parts of the Earth in

feedback loops.123 In 1970, Hutchinson produced a diagram for Scientific American depicting

“Major Cycles of the Biosphere” (for instance CO2, H2O, and N2) that showed the role of

organisms (including human organisms) in the cycling of these compounds and materials

throughout the different parts of the Earth.124 The atmospheric sciences also made use of similar

flow diagrams to depict Earth processes. For example, Roger Barry and Richard Chorley’s

popular and enduring textbook Atmosphere, Weather, and Climate (1968) contained a visual

representation of the movement of carbon throughout the Earth’s atmosphere, biosphere,

geosphere, and hydrosphere and showed how these movements determined carbon dioxide levels

in the atmosphere.125

The US Committee for GARP—a group that included Francis Bretherton among its

members—expanded on these diagrams in Understanding Climatic Change (1975), its

recommendation for a national climatic research program to improve knowledge of the basic

123 G. Evelyn Hutchinson, “Circular Causal Systems in Ecology,” Annals: New York Academy of Sciences 50, no. 4 (1948): 223. Hutchinson’s student Howard Odum built on these flow diagrams by illustrating ecological systems as analogous to electrical circuits. See: Peter J. Taylor and Ann S. Blum, “Ecosystems as Circuits: Diagrams and the Limits of Physical Analogies,” Biology and Philosophy 6 (1991): 275-94. 124 G. Evelyn Hutchinson, “The Biosphere,” Scientific American 223, no. 3 (Sep. 1970): 50-1. 125 The first edition was published in 1968 with the tenth edition appearing in 2010. Roger G. Barry and Richard J. Chorley, Atmosphere, Weather, and Climate, first edition (London: Methuen & Co., 1968), 25-6.

179

forces affecting the “global climate.” The committee pictured the Earth and its climate as “the

coupled atmosphere-ocean-ice-earth climatic system” (Figure 1.1). In a single glance, a viewer

could identify the different components of the Earth—an atmosphere with a billowy cloud, an

ocean with waves, ice, and mountainous topography, with the deeper Earth and outer space

external to the system—and how these components interacted with each other as part of an

integrated whole, as part of a system. Arrows, representing more abstract internal processes (e.g.

heat exchanges and evaporation, etc.) connect these components, or indicate external inputs to

the system (e.g. solar radiation).126 Given Bretherton’s involvement with both the US

Committee for GARP’s publication and his chairmanship of the ESSC, this diagram of the

coupled atmosphere-ocean-ice-earth climatic system is arguably a prototype for the ESSC’s

wiring diagram.

Emerging from these threads, the ESMWG’s wiring diagram of the Earth system was

abstract like the flowcharts for computer programs, but drew on the representational techniques

of Hutchinson and the US Committee for GARP to depict the major components of the Earth

system that worked on the timescale of decades to centuries. Various drafts of this diagram

appeared soon after the group’s February 1985 meeting.127 The ESMWG broke the Earth system

into two major systems (the physical climate system and the biogeochemical cycles system) that

were further broken into interconnecting subsystems with arrows linking the subsystems via flows of matter and energy. Taken as a whole, the entire wiring diagram represented what

Bretherton called a “consensus picture” of the large-scale architecture of the Earth system. The

126 United States Committee for the Global Atmospheric Research, National Research Council, Understanding Climate Change: A Program for Action (Washington, DC: National Academy of Sciences, 1975), 13-4. 127 Diagram, The Earth System: Decades to Centuries, [1985], Folder 91, Earth System Sciences Collection, NCAR, Boulder, CO.

180

processes in each of the subsystem boxes were more “glibly” assumed and required further

scientific elaboration.128

Though initially more dismissive of geology and geophysics research areas that focused on long-term processes, the ESSC and ESMWG continued discussions about timescale

prioritization and how to define the boundaries of the Earth system throughout 1985, further

evidence that ESSC and ESMWG members viewed the Earth system less as a “real” thing and

more as something pliable and indeterminate being brought into being by definition.

McCauley’s notes from the ESSC’s December 1985 meeting in San Francisco report

Bretherton’s concerns about what the focus of ESS should be. Bretherton claimed to have,

“Been haunted by the fact that we have at least 2 systems. Earth as a planet...and earth as a

system.”129 These phrases—“earth as a planet” and “earth as a system”—were shorthands that

Earth scientists used to make distinctions between the timescales for processes, with the former

referring to “planetary” processes that occurred on longer geological and geophysical timescales

while the latter refers to “system” processes that occurred on shorter, more human-centric

timescales. McCauley does not elaborate on Bretherton’s concern, but it suggests that

Bretherton viewed “planetary” processes as being outside or beyond the “Earth system” that was the focus of ESS. It might have been both practical and useful to make this distinction but, as will be shown, it caused friction between different Earth science communities.

The emphasis on the timescale of decades to centuries threatened to marginalize geological and geophysical research on and below the planet’s surface. Placing boundaries on the Earth system also threatened geophysical research on the upper layers of the Earth’s atmosphere (for instance the ionosphere or magnetosphere) and solar-terrestrial physics (aka

128 Fisher, “One Model,” 59. 129 ESSC Meeting Notes, Dec 1985, Folder 101, Earth System Sciences Collection, NCAR, Boulder, CO.

181

space plasma physics or simply space physics). The ESSC’s Working Framework (1984) placed

the upper boundary of the Earth system, “near the mesopause, at a height of about 80 km” and this marker was the upper boundary for the committee’s subsequent work.130 According to the

Working Framework, there were many important fluxes—movements of energy and matter—

between the Earth system below and the territory above that marker. For example, these fluxes

could be inputs of solar energy into the Earth system or tidal fluxes from the Earth system to the

upper atmosphere. However, the ESSC’s position was that the direct couplings here were,

“relatively weak, and it seems sensible to treat solar and space plasma physics as a separate subject [from ESS].”131 The use of the word “sensible” rather than, say, “necessary” is telling.

Paul Blanchard, in his October 1985 draft Executive Summary for the ESSC’s final report

(which became the Overview report), noted that, “As an arbitrary but practical rule, the Earth

system is taken to be contained within that layer of the atmosphere known as the mesopause,

about 85 km above the Earth’s surface.”132 Blanchard noted that processes above this were part

of the domain of “Solar-System Space Physics” and though these measurements were important

to ESS, they would serve as external inputs to the system, and not as internal, interacting

components of the system.133

130 By the publication of Closer View, the ESSC defined the upper boundary of the Earth system at the mesopause, “some 80-90 km above the Earth’s surface.” See: ESSC, Closer View, 19. 131 ESSC Working Framework, Sep 1984, ESSC/Bretherton Committee File, Box 18042, NASA HRC, Washington, DC. 132 Blanchard appears to have meant “conventional” rather than “arbitrary.” To call the Earth system boundary arbitrary implied that there were no reasons for placing its boundary where it had been placed. However, there were reasons for placing the boundary where it had been placed (e.g. not making Earth system models too complex, the different timescales for interactions, etc.) even if there could be other reasons given for placing the boundary somewhere else. It would, therefore, be more proper to call Blanchard’s Earth system boundary placement “conventional” rather than “arbitrary.” 133 Memorandum from Paul Blanchard to Earth System Sciences Committee, 1 October 1985, Folder 42, Earth System Sciences Collection, NCAR, Boulder, CO.

182

POINTS OF FRICTION

Solid Earth Science in the Margins

The Earth Systems Modeling Working Group’s definition of the Earth system focused on the shorter timescale of decades to centuries, with boundaries of the Earth system extending from the surface up to the menopause, around 85 kilometers above the ground. This disturbed many in the geology and geophysics communities (the “solid Earth community”) who grappled with longer timescales of thousands to millions of years. These were not merely disciplinary squabbles. The proposed ESS research program would be expensive and would therefore command a large proportion of Earth science research funds if it was adopted. There would be financial consequences for scientists working in areas that were marginalized by the proposed research program. Scientists conducting research on the Earth’s core or its mantle or the ionosphere or the magnetosphere would be effectively cut off from the research program and any funding. It may have been practically and conceptually useful to divide Earth science processes by time scale rather than to stick with traditional disciplinary boundaries, but the ESMWG’s decision threatened to make geologic and geophysical research priority backwaters. From a disciplinary point of view, this meant that a discipline like atmospheric science was “in” but one like geology was “out” of a proposal for a major government-funded research initiative.

Erik Conway argues that the ESSC initially intended the ESS research program to incorporate all of the Earth sciences (including geology and geophysics), though subsequent funding issues for NASA’s EOS satellites necessitated a narrowing of scope. This meant that, by

1999 and the launch of the first EOS satellite (Terra), ESS was basically synonymous with climate change science and generally focused on processes occurring on a timescale of decades

183 to centuries.134 Unlike climate change science, the version of ESS presented in the ESSC’s final report did include some “solid Earth” science, but its inclusion was more of a concession rather than the original intent of the committee. It represented the achievement made by solid Earth science representatives who fought for their research areas to be included. Despite these efforts, the ESSC’s final report still gave a lower priority to the solid Earth sciences when compared to those disciplines that focused on changes occurring on shorter timescales. Conway stresses that

NASA’s difficulties in the 1990s to maintain political support and funding for its ESS research program and EOS satellites resulted from the “gigantism” of the proposed satellites, rather than the specifics of the scientific program on offer.135 This may have been so in the 1990s, but in the

1980s, as friction with solid Earth scientists shows, the “science program” of ESS was problematic, with different groups pushing for different scientific priorities.

The fact of friction between practitioners from different Earth science disciplines is not surprising. The ESSC’s goal—to propose a research program to study the Earth as a system— was an ambitious attempt to coordinate the activities of disciplines that had long functioned autonomously and often had little in common. Interdisciplinary science has often proved controversial. In part, this results from what Conway identifies as the major role played by scientific disciplines: they, “provide the social infrastructure for the sciences: physicists determine what methodologies were acceptable in physics, established standards of evidence for physics, edited the physics journals, and operated the all-important peer review system for physics, as did geologists for geology, astronomers for astronomy, and meteorologists for meteorology.”136 These boundaries were somewhat permeable, but work that explicitly declared

134 Conway, Atmospheric Science at NASA, 301. 135 Ibid, 242. 136 Ibid, 112.

184

itself to be interdisciplinary was often viewed skeptically. Interdisciplinarity implied that

scientists were no longer bound by the traditional mechanisms for establishing scientific

authority and consensus in any specific discipline. Hyperbolically, this might imply having no

standards at all. More realistically, it meant that scientists with different backgrounds and

training needed to agree on scientific methods, standards, goals, and priorities. ESS’s vast

interdisciplinary scope required communities with different scientific methods and interests to

agree, something the ESSC arguably struggled with. As it turned out, the main sticking point

was not a specific disagreement over scientific methods and standards but a more general dispute

about how disciplinary goals and priorities would be subsumed within ESS.

According to NASA’s Shelby Tilford, in the early 1980s there was, “a real problem

between the different disciplines in Earth Science...that is oceanographers would barely talk to

atmospheric scientists, and neither one of them [would] talk to land scientists. In addition there

was the geodynamics/solid Earth community which looked at things on a completely different

time scale. There was simply very little interdisciplinary communications.”137 Tilford’s account

of the antagonisms between atmospheric scientists and oceanographers may reflect more the

internal situation at NASA rather than a deep divide between atmospheric scientists and

oceanographers.138 There were, of course, some methodological differences. For instance,

Conway notes that a number of atmospheric processes occur on much shorter timescales than

those in the oceans (compare, say, the formation of clouds that occur on scales of seconds to

hours to the circulation of ocean water that can take as long as years).139 However, James

137 Tilford interview, 23 June 2009. 138 By a number of accounts, Tilford, with a research interest in upper atmospheric research, and Stan Wilson, head of NASA’s ocean program, did not get along. Whether for personal reasons or professional competition for scarce satellite resources (at the time, Tilford was pushing UARS while Wilson backed TOPEX), it created a rift between these scientific communities at NASA. Personal correspondence. 139 Conway, Atmospheric Science at NASA, 228.

185

Fleming’s Inventing Atmospheric Science (2016) provides ample evidence of frequent collaborations between these communities, at the very least in a conceptual sense if not a practical one.140 How oceans and the atmosphere interact and affect each other has long been an

area of research interest. Some of the earliest attempts to couple numerical models from

different Earth science disciplines were attempts to link atmospheric and ocean models.141

Bretherton himself was a mathematical modeler whose research included ocean-atmosphere

interactions.142 By and large, atmospheric scientists and oceanographers had a history of some disciplinary cooperation by the 1980s.

Tilford’s recollection nevertheless highlights a basic difficulty faced by the ESSC in its attempt to bring together all the Earth sciences: the difference of timescale. The ESSC distinguished between longer and shorter timescales. On longer timescales of thousands to millions of years, the ESSC pictured the Earth as driven by two energy sources: internal radioactive and primordial sources and an external solar source. The radiative and primordial internal energy sources drove two systems. The first was the deep core-mantle system, which was dominated by the dynamo-produced magnetic field and reservoir of heat of the core, along with the physical and chemical properties of the mantle. This system interacted with the second system driven by internal energy sources, the plate tectonics system, which operated close to the

Earth’s surface and contributed to the shaping of the Earth’s topography. External energy from the sun contributes to erosion and the formation of sedimentary rocks over millions of years.143

These are research areas dominated by geology and geophysics, the solid Earth sciences. On

shorter timescales of decades to centuries, the ESSC argued that the physical climate system and

140 Fleming, Inventing Atmospheric Science. 141 ESSC, Closer View, 98. 142 For instance, see: Bretherton, “Ocean Climate Modeling.” 143 ESSC, Closer View, ch. 3.

186

the biogeochemical cycles system were most important. In the physical climate system,

atmospheric and oceanic processes govern temperature and rainfall distributions, as well as snow

and ice cover. In the biogeochemical cycles system, the flows of key nutrients such as carbon,

nitrogen, phosphorus, and sulfur influence living organisms and the chemical composition of the

land, air, and water.144 These are research areas dominated by atmospheric science,

oceanography, hydrology, and ecology.

There had been some genuine attempts by the ESSC to avert a skew against solid Earth

research. Two ESSC working groups on the solid Earth were established at the Charlottesville

summer study in 1984, one each to consider the areas of geology and geophysics. As the terms

of reference for the Geophysics Working Group explicitly stated, the group’s work was, “a

device to insure that the views of the geophysics community are properly incorporated.”145

Though there were two solid Earth working groups, this was a “practical rather than

fundamental” distinction. Geology focused more narrowly on the “outermost part of the solid

earth,” while geophysics included all parts of the solid earth and parts of solar-terrestrial physics as well.146 The aim of both groups was to safeguard the interests of these “solid Earth”

communities in the face of a program that was oriented towards disciplines that focused on

shorter term processes like atmospheric science, oceanography, and ecology. However, it was a

difficult task, given that the majority of ESSC and working group members were scientists from

those disciplines focused on shorter timescales.

The ESSC formed not one but two working groups for the solid Earth sciences.

However, these were only established after the ESSC’s first summer study in Charlottesville in

144 Ibid, ch. 4. 145 Letter from Edward A. Flinn, 20 Sep 1984, Folder 86, Earth System Sciences Collection, NCAR, Boulder, CO. 146 Draft Report of Geology Working Group for the Earth Committee (ESSC), [Jun 1985], Folder 73, Earth System Sciences Collection, NCAR, Boulder, CO.

187

June 1984, the summer study where the ESSC made crucial decisions about the prioritization of

shorter timescales and sketched out what would become its Working Framework. Nor were

Barry Raleigh (eventual chair of the Geophysics Working Group) or Kevin Burke (eventual chair

of the Geology Working Group) present at Charlottesville. This might explain the lack of

emphasis on solid Earth processes in the Working Framework, where, of the three major

components of the Earth system presented in that document (the physical climate,

biogeochemical cycles, and solid Earth systems), it was the solid Earth system that was

considered the most loosely interlinked and therefore least relevant for ESS research. Burke’s

telemail comments to Rosemary Emerson at SGC on the November 1985 draft of the Overview

noted the lack of solid Earth research: “Document [Overview] is much better. It still isolates

solid Earth as peripheral to main atmospheric and oceanic thrust. A result made inevitable by the

skewed selection of committee members two years ago. What can we do?”147 Even just

securing mention of the USGS as one of the federal agencies that could usefully participate in a

federally-coordinated ESS research program took many back and forth telemail negotiations and

reviewer comments to realize.148 Burke’s comments on the December 3, 1985 Overview draft

expressed the implications of this lacuna in no uncertain terms: “Putting the USGS in a

subordinate role reveals the extent to which geology and geophysics are regarded in this draft as

trivial compared to atmospheric and ocean sciences.”149

147 Telemail from K. Burke to R. Emerson, 18 Nov 1985, Folder 37, Earth System Sciences Collection, NCAR, Boulder, CO. 148 Telemail from L. McCauley to F. Bretherton, 27 Sep 1985, Folder 133, Earth System Sciences Collection, NCAR, Boulder, CO; Note from Kevin [Burke] to Francis [Bretherton] and Paul [Blanchard], 13 Dec 1985, Folder 133, Earth System Sciences Collection, NCAR, Boulder, CO; Telemail from E. Flinn to L. McCauley, 3 Mar 1986, Folder 12, Earth System Sciences Collection, NCAR, Boulder, CO. 149 Telemail from L. McCauley to F. Bretherton, 27 Sep 1985, Folder 133, Earth System Sciences Collection, NCAR, Boulder, CO.

188

Despite attempts to reconcile the divergent interests of different Earth science

communities, the ESSC viewed it as a matter of practical necessity to focus on global changes

occurring on shorter timescales. In January 1985, the ESSC held its fifth official meeting at JPL,

though it was really the first full meeting of all members. Berrien Moore, a biogeochemical

cycle expert (and not a solid Earth scientist), noted the, “need to divide Earth System Sciences by

timescales” and then pick something specific to prioritize. The meeting concluded in part by

listing the, “two alternatives for treatment of Solid Earth [science] in the final report[.]” First,

that the ESSC do what Moore suggested: divide ESS research by timescale and give the various

timescales separate flagship missions. Two, there was a “Strongly held minority view that [there

was] only one subject, the earth system sciences, and that we [should] treat solid earth within that

framework.”150 Which of these alternatives was preferable depended on whether an ESSC

member believed that Earth scientists researching processes of vastly different timescales could

conduct interdisciplinary research in a coordinated fashion and could practically build a

numerical model of the Earth system that included all of these processes.

There were those outside the solid Earth community who expressly encouraged

constructing an ESS research program that focused only on shorter timescales. Preparations of

the Closer View included the compilation of tables that ranked measureable global variables in the Earth system according to their importance (see below). Atmospheric physicist Bruce

Barkstrom—who worked on NASA’s ERBE—was one of the scientists Bretherton asked to review these tables in early January 1987.151 Barkstrom acknowledged that the tables were, as

Bretherton had already noted, “a tradeoff, involving numerous considerations” and hinged on

150 General Notes, Fifth Meeting of the ESSC, Pasadena, CA, [Jan 1985], Folder 117, Earth System Sciences Collection, NCAR, Boulder, CO. 151 Memorandum from F.P. Bretherton to J. Carey et al., 2 Jan 1987, Folder 149, Earth System Sciences Collection, NCAR, Boulder, CO.

189

what was considered important for ESS and studying global change. For Barkstrom, the choice

was obvious: “it would seem that the measurements of solid earth properties, notably mantle

structure and plate deformations, belong on a list with a much longer time scale. While these are

undoubtedly important, their direct impact on our lives would appear to be small in comparison

with the potential effects of either a major reduction in O3 or with the expected changes in

climate owing to increased trace gases.”152 Solar astronomer Jack Eddy agreed. In his

comments on the November 1985 Overview draft, Eddy took issue with the prioritization of

geological measurements like plate tectonic motion, continental deformation, mantle circulation,

and magnetic field generation over other measurements: “wait a minute! Why are these more

important in terms of practical benefits than say studies of SOILS or VULCANISM or

EARTHQUAKES. This reads like the geology lobby to me.”153 In his comments on a draft of

the Closer View from May 3, 1987, John Dutton begged the ESSC not to start off the section

focused on Global Changes: Decades to Centuries “by going back to millions of years” since the

longer timescale “has too much emphasis now.”154

Many in the solid Earth community disagreed. As late as May 1986 (the month the

Overview report was published) there were still protests against the marginalization of solid

Earth science. McCauley, writing via telemail to the ESSC members regarding the ESSC exhibit

at the Spring American Geophysical Union meeting, reported on a conversation with Raymond

Arvidson, a planetary geologist. Arvidson was, “very unhappy with our program and how we

have blatantly left out the solid earth - as demonstrated by our [the ESSC’s] wiring diagram. He

152 Telemail from B.R. Barkstrom to L. McCauley, 18 Jan 1987, Folder 141, Earth System Sciences Collection, NCAR, Boulder, CO. 153 Earth System Sciences Overview 86, Draft with Comments, 7 Nov 1985, Folder 37, Earth System Sciences Collection, NCAR, Boulder, CO. 154 Telemail from J. Dutton to the ESSC Executive Committee, 3 May 1987, Folder 136, Earth System Sciences Collection, NCAR, Boulder, CO.

190

mentioned Raleigh and Burke’s continued unhappiness that the committee has not listened or

understood the importance of solid earth science. [Arvidson] says the program is unbalanced

and that indeed he has spoken with FPB [Bretherton] and he just doesn’t listen.”155 This

exchange occurred after the ESSC’s Overview report had already gone to the publisher for

printing (it was published in May 1986), and barely a month before its press conference to

publicly unveil the ESS research program (chapter four). Burke raised these concerns back in

December 1985 in a review of the Overview draft, noting that the original wiring diagram,

“omits geology and geophysics almost completely….It is not easy to see how this is repaired,

since there is nothing in the figure on the solid earth.”156 By the time of the Closer View report,

Burke had remedied the situation by constructing a second wiring diagram to include global

changes that occurred on timescales of thousands to millions of years, processes that were

neglected in the original Earth system diagram developed by the ESMWG (Figure 3.2).157

However, these processes were still ranked lower in importance than shorter term processes,

even if now they had their own wiring diagram.

155 Telemail from L. McCauley to the ESSC Executive Committee, 21 May 1986, Folder 137, Earth System Sciences Collection, NCAR, Boulder, CO. The telemail text has been corrected for spelling mistakes. 156 Telemail from L. McCauley to F. Bretherton, 27 Sep 1985, Folder 133, Earth System Sciences Collection, NCAR, Boulder, CO. 157 ESSC, Closer View, 26.

191

Figure 3.2. Earth system wiring diagram from the ESSC’s Closer View (1988), depicting Earth processes occurring on timescales of thousands to millions of years. (Design by InterNetwork, Inc | Payson R. Stevens; reprinted with permission from NASA)

According to Geophysics Working Group chair Raleigh, the deprioritization of solid

Earth science research was not limited to the global variable observation tables or the Earth system wiring diagram but infused the entire final ESSC report. On May 11, 1987, Bretherton wrote to the ESSC executive committee to inform them of the status of the Closer View and its reviewer comments. Included was a comment from Raleigh suggesting additional sentences be

192

added to the prologue to explain that geophysics research was being given less emphasis than

processes in the Earth’s oceans, atmosphere, and biota because changes in the latter areas were

occurring more rapidly. The reason was emphatically not because the Earth’s interior was less

scientifically important or interesting. Bretherton and the other core ESSC members rejected the

suggested revision on the grounds that this was inappropriate in the prologue and that the issue

had been addressed in other parts of the report.158

Ultimately the two factions came to enough agreement about the final text of the Closer

View report that it could be published in January 1988. The interim compromise in the Overview

was to make the “goal” of ESS as broad and vague as possible so that it could, at least in

principle, incorporate all of the Earth science disciplines. The nebulous goal of ESS was, “To

obtain a scientific understanding of the entire Earth System on a global scale by describing how

its component parts and their interactions have evolved, and how they may be expected to

continue to evolve on all timescales.”159 In its broadest sense, ESS was focused on the entire

Earth system for all timescales. However, the ESSC immediately narrowed the focus of ESS to

the preferred timescale of decades to centuries. The narrower ESS “challenge,” in contrast to the

“goal,” was, “To develop the capability to predict those changes that will occur in the next

decade to century, both naturally and in response to human activity.”160 The Closer View

reiterates the “goal” and “challenge” as characterized by the Overview.161 An entire chapter in

the Closer View is devoted to Earth processes occurring on longer timescales of hundreds to

millions of years, but overall priority is still given to those processes occurring on shorter

timescales. A single chapter, along with the additional wiring diagram depicting the Earth

158 Telemail from F. Bretherton to the ESSC Executive Committee, 11 May 1987, Folder 116, Earth System Sciences Collection, NCAR, Boulder, CO. 159 ESSC, Overview, 4 [emphasis added]. 160 Ibid [emphasis added]. 161 ESSC, Closer View, preface.

193

system on longer timescales, was the extent of the concessions the solid Earth community could

obtain for their research within the ESS program. As a result, many in the geophysics and

geology communities viewed with ambivalence the ESS research program that effectively

deprioritized their work (chapter five).

One Implementation Strategy To Rule Them All

Another major point of contention for the ESSC was also one of the main reasons for its

existence: the development of an implementation strategy for the ESS research program.

Perhaps Earth scientists could broadly agree that the Earth should be thought of as a system with

interconnecting subsystem components, and that it was increasingly important, as well as

practically possible, to study these components together rather than in isolation. But how would

one go about actually studying the Earth in such a manner? What would the specifics of an Earth

system science research program look like? What were the major scientific issues to consider

and how should they be prioritized? What kinds of measurements were needed and what kinds

of instruments and techniques—space-based or otherwise—would best achieve these goals?

These were the kinds of questions that an ESS implementation strategy could answer, questions

that had not been addressed by any previous Earth science committee or report.162 Bretherton

stated to ESSC members that its outputs must provide details on costs and scientific priorities: “it

is quite clear that a forthright attack on these is mandatory if our report is to avoid being

consigned immediately to the trash can.”163

162 Ibid. 163 Earth System Sciences Committee, Status Report on Major Issues, Francis P. Bretherton, [1985], Folder 73, Earth System Sciences Collection, NCAR, Boulder, CO.

194

An implementation strategy was important for scientific reasons, but it also carried

significant financial implications. Those Earth sciences that received higher priority would,

presumably, receive more funding for their research. The result was a constant tension between

the desires of scientists from each discipline and the overarching need to set priorities. As

Shelby Tilford expressed it, “Everybody wanted to be first. Well, we didn’t want everybody to

be first. We wanted everybody to work together. This was hard.”164 Any implementation

strategy ran the risk of creating an Earth science discipline hierarchy, with the haves at the top

and the have-lesses and have-nots further down. When the stakes for research funding were

potentially so high, it is hardly surprising that negotiations to develop an ESS implementation

strategy would be intense and protracted. It also helps explain why Earth system science as a

specific research program with specific scientific priorities was received by the Earth science

community with ambivalence, while the vaguer concept of understanding the Earth as a system—the Earth system—was met with some enthusiasm and spread widely (chapter five).

On its surface, the ESSC’s recommended implementation strategy in the Closer View report represented a balanced program for ESS. It contained seven crucial components: (1) sustained, long-term measurements of global variables on all timescales; (2) a fundamental description of the Earth and its history via measurements that, once obtained, did not require repetition unless improved instruments became available;165 (3) research foci and process studies

that focused on key, though not necessarily global, Earth science problems; (4) the development

of Earth system models; (5) the development of an information system to facilitate analysis and

numerical modeling; (6) the coordination of US federal agencies with respect to funding and

164 Tilford interview, 23 June 2009. 165 These observations included land-surface data like “topography” and geophysical measurements for “gravity and geoid” and “crustal magnetism.” See: ESSC, Closer View, 150-1.

195

organizing ESS research; and, (7) international cooperation. The program called for a two-

pronged strategy with near term and long-term objectives. For the near-term period of 1987-

1995, the ESSC recommended a number of strategies that took advantage of already active or planned projects in line with ESS objectives. It recommended the enhancement and continuation of operational Earth observations like NOAA’s series of Geostationary Operational

Environmental Satellites (GOES) that had collected meteorological data from geostationary orbit since GOES-1 launched in 1975. It also recommended the “timely completion” of more specialized satellites already in development, like ERBE, UARS, and TOPEX. The ESSC suggested that NASA develop an Earth System Explorer series of satellites that would include the Geopotential Research Explorer Mission (GREM). Other operational instruments should be flown on suitable platforms to obtain required measurements, like an ocean colour scanner and an atmospheric carbon monoxide monitor. The ESSC recommended that current basic science research and in-situ studies undertaken by agencies like NASA, NOAA, NSF, the USGS, the

Department of Energy (DOE), and the Office of Naval Research be expanded and coordinated.

An information system would need to be developed that could store and process global datasets, facilitating data analysis and the development of numerical models of the Earth system. New instruments and techniques should also be developed in the near term in preparation for future satellite missions. In the near term, the ESSC envisioned ESS as assembling and integrating current research efforts to develop a more coherent and coordinated approach to achieving ESS objectives. For the era of 1995 and beyond, the ESSC recommended focusing on a new observational system, NASA’s EOS, comprised of a series of polar-orbiting platforms containing

196

numerous instruments on a single platform to continue the collection of integrated global

observations that had begun during the near-term implementation phase.166

These implementation details were coalesced only after extensive debate. At an early

ESSC meeting in January 1985, this strategy received “intensive discussion.”167 In preparation

for the second summer study in June 1985, the ESSC’s executive committee pulled together a

rough draft of its report, divided into 5 chapters: (1) broad scientific rationale, (2) essential

science, (3) observing strategies, (4) implementation strategies and priorities, and (5) agency

roles. The implementation section was still in a sketchy state when compared to the other

chapters, with only point form notes and no scientific prioritizations.168 Minutes of a full ESSC

meeting in September 1985 singled out the implementation strategy chapter as being, “the least

well developed” and indicated that the scientific priorities required further review and

agreement. In fact, it was the incomplete status of this chapter that delayed the publication of the

ESSC’s final report.169 In early 1987, Bretherton also conducted meetings with external

representatives from NOAA, NSF, and NCAR to further discuss any “issues” with the tables

developed by the ESSC listing the required scientific measurements for global variables required

for ESS.170

One issue that emerged during these discussions was how much weight to give to the

EOS platform at the expense of smaller, more targeted space systems. The proposed EOS would

be a series of large, polar-orbiting platforms, larger than any Earth observing satellites yet

166 ESSC, Closer View, 134-7. 167 Earth System Sciences Committee, Status Report on Major Issues, Francis P. Bretherton, [1985], Folder 73, Earth System Sciences Collection, NCAR, Boulder, CO. 168 Outline for ESSC Final Report, [Mar 1985], Folder 73, Earth System Sciences Collection, NCAR, Boulder, CO. 169 ESSC Update No. 4, [1 Oct 1985], Folder 71, Earth System Sciences Collection, NCAR, Boulder, CO. 170 Memorandum from F.P. Bretherton to J. Carey et al., 2 Jan 1987, Folder 149, Earth System Sciences Collection, NCAR, Boulder, CO.

197

built.171 Each platform could carry a number of instruments that could take simultaneous

measurements for a number of different variables. Grand in design, the EOS would be expensive

and potentially dangerous to fly, since so many instruments on a single platform would make a

host of needed measurements susceptible to a single launch failure. Cost estimates varied

throughout the ESSC’s discussions, but they were always high, calling for a substantial funding

increase for NASA’s Earth observing satellite budget. As Conway notes, when Congress

approved funding for EOS in 1989, the program became “the most expensive science program in

American history,”172 with costs expected to reach around $17 billion by the end of fiscal year

2000.173 The actual development and approval of EOS (and its eventual redesigns, rescopings,

rebaselinings, and restructurings as part of NASA’s Mission to Planet Earth) is beyond the scope

of this dissertation and is covered extensively by Conway, Goldstein, Lambright, McElroy and

Williamson, and Leshner and Hogan. What is important here is that the ESSC extensively

debated whether and how much to link its work to EOS, and thus how much of its

implementation strategy would center around EOS. If too much emphasis was placed on the

EOS platforms, then that would come at the expense of smaller, less expensive satellites that

might carry only a single instrument targeting a specific variable. It would also mean more

compromises and tradeoffs between the kinds of science that could be done, and what kinds of

instruments could practically and usefully be placed on a single platform in a single orbit.

171 NASA’s two initial platform conceptions were each to weigh around 12,210 kg and carry 24 instruments divided between the two platforms. For comparison, the next largest satellite ever conceived, the Upper Atmosphere Research Satellite (UARS), was to weigh 6,736 kg. Dimension-wise, one of the proposed EOS platforms was to be 4.3 meters in diameter, and 12 meters high. UARS was to be 4.3 meters in diameter and 9.8 meters high. See: NASA, From Pattern to Process, 20; Richard A. Kerr, “Why Bigger Isn’t Better in Earth Observation,” Science 253, no. 5027 (27 Sep. 1991): 1481; Lambright, “The Political Construction of Space Satellite Technology,” 63. 172 Conway, Atmospheric Science at NASA, 199. Conway does not specify what is included under the phrase “science program.” 173 United States Senate, NASA’s Space Science Programs and the Mission to Planet Earth: Hearing before the Subcommittee on Science, Technology, and Space of the Committee on Commerce, Science, and Transportation, United States Senate, One Hundred Second Congress, First Session, April 24, 1991 (Washington, DC: GPO, 1991), 4.

198

Some ESSC members and external reviewers questioned the wisdom of aligning ESS too

closely with EOS. Bretherton raised the issue in his list of discussion points for the ESSC’s

January 1985 meeting, noting that the EOS would be an expensive pursuit and would require a strong scientific rationale to make it fiscally feasible. Notes from this meeting reported that

Bretherton raised this concern, stating that the ESSC must look more broadly for scientific rationales for ESS than simply the “EOS report[.]” Later in the discussions, engineer Vic Reis wondered if the EOS report was the “gut of ESS recommendation?”174 Perhaps fearful that the

EOS’s high costs would make the ESS research program vulnerable, soil and vegetation specialist Paul Zinke argued during the December 1985 ESSC meeting that ESS should have a broader implementation strategy than simply the EOS platforms.175

Since the EOS platforms promised to simultaneously collect more data on more variables than had ever been collected before, the appeal of these platforms for ESS is readily apparent.

NASA officials also had a vested interest in having the ESSC’s work align with EOS, as a scientific and programmatic justification for the satellite system. Copies of reports by NASA’s

Science Steering Committee for EOS were made widely available to ESSC members. Dixon

Butler, Shelby Tilford’s second in command at the OSSA’s Division of Earth Sciences, gave a special presentation on EOS to ESSC members on April 11, 1985.176 Butler also presented at the

ESSC’s second summer study in June 1985, where Butler described the ESS research program as

an “EOS enabler.”177 Another document from this summer study detailing the “Vulnerabilities and Dangers in Establishing this [ESS] Program” stated that “EOS is the centerpiece of the ESS

174 General Notes, Fifth Meeting of the ESSC, Pasadena, CA, [Jan 1985], Folder 117, Earth System Sciences Collection, NCAR, Boulder, CO. 175 ESSC Meeting Notes, Dec 1985, Folder 101, Earth System Sciences Collection, NCAR, Boulder, CO. 176 Telemail from L. McCauley to K. Wolfe, 13 Nov 1984, Folder 86, Earth System Sciences Collection, NCAR, Boulder, CO; Telemail from L. McCauley to D. Butler, 2 Apr 1985, Folder 77, Earth System Sciences Collection, NCAR, Boulder, CO. 177 [Dixon] Butler’s Vugraphs, [Jun 1985], Folder 73, Earth System Sciences Collection, NCAR, Boulder, CO.

199

program[.]”178 Bretherton himself gave mixed signals about the relationship between ESS and

EOS. Documents prepared for the second summer study in June 1985 by McCauley and

Bretherton explicitly stated that, “The ESSC must come to own EOS. Therefore, we need to

review EOS carefully.”179 This position differs from the one Bretherton articulated only months

before, when he expressed reservations about aligning ESS too closely with EOS. After

reviewing an ESSC draft report in September 1985, James McCarthy argued that it did not

adequately justify the requirement for EOS and, “why we need all this good stuff flying at the

same time on the same birds.” This was an oversight if, as McCarthy believed, the ESSC, “was

to build the case that requires EOS, and hence the polar platforms[.]”180 At the end of 1986,

Berrien Moore worried that the ESSC was “dangerously near a fatal mistake” by not fully and clearly endorsing EOS. The ESSC, according to Moore, must promote the EOS, “as the preferred means of implementing the needed [ESS] measurement program,” which requires comprehensive, integrated, simultaneous, global observations.181 Ultimately, the EOS would

serve, according to the Closer View report, as “the polar-orbiting centerpiece of a comprehensive observing program for Earth system science.”182

Given the emphasis placed on EOS, it is not surprising that many scientists viewed the

ESS research program as too heavily prioritizing space-based measurements and instruments over in-situ ones. Remote sensors—be they passively collecting incoming electromagnetic radiation or actively firing signals and measuring the resultant effects—generally cannot

178 Vulnerabilities and Dangers in Establishing this [ESS] Program, [1985], Folder 23, Earth System Sciences Collection, NCAR, Boulder, CO. 179 Issues, [20 Mar 1985], Folder 73, Earth System Sciences Collection, NCAR, Boulder, CO. 180 Telemail from L. McCauley to F. Bretherton, 27 Sep 1985, Folder 133, Earth System Sciences Collection, NCAR, Boulder, CO. 181 Telemail from B. Moore to the ESSC Executive Committee, 31 Dec 1986, Folder 156, Earth System Sciences Collection, NCAR, Boulder, CO. 182 ESSC, Closer View, 138.

200

penetrate the Earth’s land or water surfaces. This means that they are largely limited to

observing atmospheric conditions and surface features. For disciplines that examine subsurface

processes, whether on land or in water, remote sensing must be augmented with robust in-situ

observations. Balance between space and non-space data was considered important, yet the

ESSC’s work did not always reflect this, and was at times decidedly space-centric. The Closer

View report continually stressed that balance was needed between satellite and in-situ

measurements, but the in-situ component of ESS mostly focused on providing data to properly

calibrate and validate satellite data. In the final version of its implementation strategy, the ESSC

did recommend an ESS research program that included in-situ process studies, but it was a much

smaller component than the proposal for satellite remote sensing observations on global

variables. Perhaps this emphasis was due to the ESSC’s origins in NASA’s Advisory Council.

After all, it had been charged with determining NASA’s role in a major ESS research program

which presumably would draw on the agency’s spaceflight expertise. It may have been

overdetermined that a group of researchers assembled to recommend an implementation strategy

for ESS for NASA (and later other federal agencies) would place heavy emphasis on

measurements from satellites.

While the outcome may not have been unexpected, it exacerbated tensions between the

solid Earth scientists and those working in disciplines that could more readily utilize data from

space-based platforms, disciplines investigating the atmosphere and the surface of water and

land. Etienne Benson describes the resulting tensions within geological communities—

specifically geomorphology—in the 1980s over whether or not “mega-geomorphology,” which

focused on larger spatial scales and utilized remote sensing data, was a legitimate avenue of

201

study for a discipline traditionally reliant on local field work and first-hand observations.183

These tensions also materialized in responses to the ESSC reports. Bretherton’s files are filled with letters from non-ESSC scientists critical of aspects of the ESSC’s activities. One

correspondent, a geology professor emeritus from the University of Missouri-Columbia, noted that the Overview report, “was heavy on instrumentation and space, at the expense of

GEOLOGY, both the word and the rocks(!).”184 Similarly, Raymond Watts from the USGS

made this view known in his review of the Overview report’s December 1985 draft. Watts took issue when the draft suggested that measurements of the Earth system are, “only available from carefully designed space-based systems[.]” In the margins he wrote, “BULLSHIT!” at this

assertion. As he more calmly noted, much useful data for understanding Earth’s climatic history

came from non-space observations, for instance the polar ice core records that showed the

influence of volcanic eruptions on the climate. Watts bluntly called for a restructuring of the

report, “to emphasize a balanced, coordinated, space and in-situ observation program, rather than

the present ‘space-first, in-situ also-ran’ structure.”185 In his revisions, Jack Eddy also noted that

an early Overview draft gave disproportionate attention to satellite observations at the expense of

in-situ research. The in-situ section of the report was, according to Eddy, “awfully brief and out

of balance with respect to the long satellite section that just preceded it.” Eddy stressed that in-

situ research was not simply important as a way to calibrate and validate satellite data, as the

report suggested, but can often be, “more fundamental, certainly to an ecologist or soil scientist

183 Etienne S. Benson, “Re-Situating Fieldwork and Re-Narrating Disciplinary History in Global Mega- Geomorphology,” Studies in History and Philosophy of Science Part A 70 (2018): 28-37. 184 Letter from Walter D. Keller to Laura L. McCauley, 24 Oct 1986, Folder 76, Earth System Sciences Collection, NCAR, Boulder, CO. 185 Comments on Overview ‘86 by Ray Watts, USGS, 3 Dec 1985, Folder 137, Earth System Sciences Collection, NCAR, Boulder, CO.

202

or almost any other.”186 While the ESSC drafted a final report that pushed ESS’s measurement

needs into alignment with NASA’s spaceflight capabilities, some scientists independent of

NASA were skeptical of the approach. For some scientists, in-situ measurements had served many disciplines well, and diverting scarce research funding to space-based platforms might not be the most effective use of research dollars.

Perhaps not surprisingly, the ESSC’s implementation strategy prioritized space-based

measurements for Earth processes occurring on the timescale of decades to centuries, with much

less concern for long-term “solid Earth” processes. This short-term emphasis was particularly

obvious in the compiled tables that listed the global variables that ESS would collect

measurements for on an ongoing basis. Members viewed these measurement priority tables as,

“the cornerstone of the implementation plan.”187 However, the ESSC recognized and admitted

that the compilation of all the desired variables, “into the confines of a single table inevitably

entails a number of oversimplifications and personal judgements; the table entries should

therefore be considered starting points for future elaboration and refinement.”188 The committee

identified 56 key variables grouped into six broad categories: external forcings, concentrations of

chemically and radiatively important trace species, atmospheric response variables, land-surface

properties, ocean variables, and geophysical variables. In its evaluation of the relative scientific

importance of each variable, the ESSC gave a rating of stars, with three stars representing the

most “essential” variables, with one star the least significant. These tables clearly prioritized

both space-based measurements and variables occurring on shorter-term timescales. Well over

half of these identified variables—34 of the 56 variables ranked—related to research topics

186 Earth System Sciences Overview 86, Draft with Comments, 7 Nov 1985, Folder 37, Earth System Sciences Collection, NCAR, Boulder, CO. 187 Memorandum from Laura Lee McCauley to Nancy Ann Brewster et al., [1987], Folder 149, Earth System Sciences Collection, NCAR, Boulder, CO. 188 ESSC, Closer View, 140.

203

focused on the atmosphere, oceans, and biota. Just five variables were identified as specifically

“geophysical.” Of the 15 variables assigned as being of “essential” importance, only one of

these was a geophysical variable (plate deformations, small scale).189 Only eight of the 56

variables could not be measured, at least in part, via Earth observing satellite instruments. Only

three of the “essential” variables could not be measured using Earth observing satellite

190 instruments (CO2, surface air temperature, and subsurface sea circulation). This selection of

emphasis could have financial implications for disciplines that needed more than surface or

atmospheric data. If Congress chose to throw its full support and financial weight behind the

ESS research program that favoured space platforms, this could then leave less funding for those

in-situ communities. Though the ESSC ultimately signed off on the ESS implementation

strategy contained in the Closer View, it was the most problematic aspect of their work to

complete, generating the most friction with external communities.

CONCLUSION

This discussion of Bretherton’s previously unexamined papers shows the full range of the

ESSC’s work in formulating a large-scale Earth science research program that treated the whole

Earth as a scientific object. With the spectre of Global Habitability’s failure ever-present, the

ESSC was motivated to correct previous mistakes by ensuring its ESS program was meticulously developed. While the specific research program was never fully adopted (chapter five), the

ESSC’s description of the “Earth system” that broadly described an interconnected planet amenable to interdisciplinary collaborative research proved particularly fruitful and long lasting

189 The other 14 “essential” variables were: , CO2, stratospheric O3, surface air temperature, tropospheric temperature, pressure (surface), precipitation, index of vegetation cover, sea-surface temperature, ocean wind stress, subsurface sea circulation, and ocean chlorophyll. 190 ESSC, Closer View, 142-9.

204 for Earth scientists. The scope of this work, combined with the ESSC’s promotional activities

(chapter four) help explain the ubiquity of the Earth system concept in the Earth sciences today.

Whatever its initial panoramic ambitions, the ESSC recognized that not all aspects of the planet could be incorporated into a practical ESS research program. The “challenge” of the

ESSC’s work was to devise ESS so that a numerical computational model of the “Earth system” might be constructed. To meet the “challenge,” prioritizations would need to be established.

The ESSC chose the human timescale of decades to centuries as being the most important, given the many environmental problems facing policy-makers and also, no doubt, due to the disciplinary composition of the ESSC (mostly scientists from disciplines that focused on shorter- term processes like atmospheric science, oceanography, and ecology). Members understood the

“Earth system” as a construct that could be defined in a number of ways, depending on specific research interests and concerns. Indeed, the Dynos joke suggests that ESSC members were aware of the extent to which the scope of their project was intimately linked to the scale of human interests embedded in the limit of a human (as opposed to a dynaperson) lifespan.

Whatever ESS would be, it was essentially and ineluctably anthropic. The ESMWG decided to define the Earth system using the prioritized decades to centuries timescale over the objections of geologists and geophysicists who favoured longer timescales. The boundaries of the Earth system were drawn in such a way that many who worked on research areas that focused on longer term processes—like those in the Earth’s core and mantle and in areas beyond the mesopause—were concerned about being marginalized in the proposed ESS research program.

Two “frictions” arose during the development of the ESS research program: friction with the solid Earth sciences community and frictions with the development of the ESS implementation strategy. These issues were intimately related. Solid Earth scientists examined long timescales of thousands to millions of years. It was, for all practical purposes, impossible to

205

reconcile these with shorter term processes into a single numerical computer model of the Earth

system. While both the long-term and short-term timescales could in the abstract make equal claim to being indispensable parts of the Earth—and both received their own conceptual Earth system wiring diagram—there was little reasonable prospect for them co-existing in a model that addressed ESS’s “challenge.” Solid Earth scientists also studied processes that could not, for the most part, be investigated with the surface or atmospheric satellite measurements prioritized in the ESS implementation strategy. Most ESSC members—coming mainly from atmospheric science, oceanography, and ecology—conducted research on the shorter timescales more amenable to satellite observations.

These issues would contribute to ESS’s lukewarm response from external scientific communities in the US. To fully understand how the Earth system concept became so ubiquitous, despite the resistance the ESS research program encountered, it is necessary to examine in detail the ESSC’s extensive promotional efforts (chapter four). Though this program did not receive the widespread support the ESSC hoped for, those exposed to its ideas readily adopted its scientific conception of the Earth as a system (chapter five).

Chapter 4 The Earth System Sciences Committee: Promoting a Research Program

INTRODUCTION

Despite all of the debates, frustrations, and delays, the ESSC did manage to produce enough

consensus to publish two reports, the Overview and the Closer View that presented a detailed

rationale for its Earth system science research program. ESS was a grand framework for

bringing together practitioners from across the spectrum of the Earth sciences, and would require

advanced and expensive technologies. It would rely heavily on Earth observing satellites, the

development of advanced numerical models, and a sophisticated information system capable of

storing and processing vast amounts of data. It would not be cheap, and would require the

coordination of expertise not just from the sciences, but also from engineering communities.

This research program would, therefore, require widespread scientific, engineering, and political

support. While many of those involved with the ESSC might have taken issue with this or that

detail, there was almost always unanimous agreement that the ESS documents and their

supporting material needed to be written and presented so that they would appeal to diverse constituencies.

These constituencies were many and varied. There were, of course, individuals from the different Earth science disciplines: atmospheric physicists and chemists, aeronomists, marine biologists, ocean dynamicists, hydrologists, limnologists, biologists, geochemists, biogeochemists, systems ecologists, population ecologists, pedologists, geophysicists, geologists, geodesists, stratigraphers, meteorologists, volcanologists, and paleoclimatologists. Though all

focused on some aspect of the planet, there were a number of issues—notably the different

timescales that practitioners worked in—that made collaborative efforts challenging. Then there

206 207

were the engineers, with specialties ranging from signal processing and optics to electronics and

radiometry.

Beyond the scientists and engineers, the ESSC needed its material to be accessible to

those in the political realm. Those reading the ESSC reports could be civil servants working for

NASA, NOAA, NSF, USGS, the Department of Agriculture, DOE, the Office of Naval

Research, the Environmental Protection Agency, the Agency for International Development,

Housing and Urban Development, the Department of State, or even the Department of Defense

(DOD).1 They could also be the political appointees that headed those governmental administrations and departments. They might be the congressional staffers forever being courted through both formal channels and informal backroom deals. Or they might be the politicians themselves who sit on important congressional committees responsible for authorizing and appropriating government discretionary spending. And this is just the national context. An important part of the ESS research program would include international cooperation, whether it

be as part of major international scientific programs like the IGBP or research projects

undertaken by the WCRP, or smaller bilateral agreements to coordinate the design and

construction of satellite platforms and instruments or the use of data.2 Then there was the general public to consider. Public support could translate into political support, and therefore financial support. Members of the public might also become future practitioners, so establishing and maintaining a large support base was important. All of these groups required consideration

1 This is the list of US federal agencies that conduct or use Earth science research compiled by the ESSC. See: Earth System Sciences Committee (ESSC), Earth System Science: A Program For Global Change: A Closer View (Washington, DC: NASA, 1988), 161. 2 Some of the WCRP’s projects in the 1980s included: the Tropical Ocean Global Atmosphere program (TOGA), the World Ocean Circulation Experiment (WOCE), and the International Satellite Cloud Climatology Project (ISCCP). See: ESSC, Closer View, 154-5.

208

as the ESSC developed its research program and worked out how to present its material to an

enormous variety of potential readers and audiences.

This chapter examines the steps the ESSC took to ensure that ESS material was

scientifically sound, accessible, and widely distributed. The group disseminated and promoted

their ideas through a variety of mechanisms and strategies. These efforts illustrate that members

were keenly aware of the importance of consensus building to develop and promote a new

scientific research framework intended to be wide-ranging in its application. This chapter, therefore, broadly aligns with other scholarship in the history of science that emphasizes the

promotional aspects of scientific activity and the work required to build and expand support

networks.3 It also fits with studies of how knowledge and ideas coalesce and are then

disseminated, and the extraordinary effort often required for this to happen. As John Krige notes

in his edited volume How Knowledge Moves (2019), the movement of knowledge is a “social

accomplishment” that is itself an achievement, and so its spread cannot be simply presumed.4

Indeed, it is now an uncontroversial claim to suggest that a crucial part of scientific

practice involves consensus building. This can even be seen in the literature that touches on the

ESSC. Work by Erik Conway, Eric Goldstein, W. Henry Lambright, Richard Leshner and Thor

Hogan, and John McElroy and Ray Williamson all look at the efforts of NASA officials and

affiliated researchers to establish a robust Earth science research program at NASA and achieve

3 Most notable here is Bruno Latour’s work on actor-network theory (ANT), in which he shows how human and nonhuman network alignments are necessary for the successful development of scientific theories and technologies. Failures occur when the associations in these networks fail to coalesce, as Latour memorably demonstrates in his study of the proposed Aramis personal rapid transit system in Paris. See: Bruno Latour, Science in Action: How to Follow Scientists and Engineers Through Society (Cambridge, MA: Harvard University Press, 1987); Bruno Latour, The Pasteurization of France, trans. Alan Sheridan, John Law (Cambridge, MA: Harvard University Press, 1988); Bruno Latour, Aramis, or the Love of Technology, trans. Catherine Porter (Cambridge, MA: Harvard University Press, 1996). 4 John Krige, “Introduction: Writing the Transnational History of Science and Technology,” in How Knowledge Moves: Writing the Transnational History of Science and Technology, ed. John Krige (Chicago: University of Chicago Press, 2019), 5.

209

support for the construction of the EOS. Doing this required many negotiations and tradeoffs

between scientists, engineers, and politicians.5 However, no one has yet looked specifically at the work done by ESSC members to construct and promote the idea of studying the Earth as a system of interconnected components. Understanding their work is key to understanding how the Earth system concept became ubiquitous in the Earth sciences. Just as the Earth system required “assembly” (chapter three), so too did an external base of supporters and practitioners for the ESS research program.

Despite the ESSC’s considerable efforts, however, success was not total or lasting. The

ESSC’s specific ESS research program proposal focused on global change occurring on timescales of decades to centuries and relied heavily on a fleet of Earth observing satellites. In

so doing, it prioritized certain kinds of research and data collection methods over others. Given

the scale of the proposed program and the prospect of being marginalized or effectively cut off

from possible future funding, many scientific communities took issue with the ESSC’s

recommended program (chapter three). Ultimately, the result would be that the specific ESS

research program developed by the ESSC failed to generate a strong and enduring support base

within scientific or political communities. Nevertheless, the general idea of studying the Earth as

an interconnected system—denoted by the phrase “Earth system”—would help coalesce the

5 Erik M. Conway, Atmospheric Science at NASA: A History (Baltimore: Johns Hopkins University Press, 2008); Erik M. Conway, “Bringing NASA Back to Earth: A Search for Relevance during the Cold War,” in Science and Technology in the Global Cold War, eds. Naomi Oreskes and John Krige (Cambridge, MA: MIT Press, 2014), 251– 72; Edward S. Goldstein, “NASA’s Earth Science Program: The Bureaucratic Struggles of the Space Agency’s Mission to Planet Earth” (PhD dissertation, George Washington University, 2007); W. Henry Lambright, “Administrative Entrepreneurship and Space Technology: The Ups and Downs of ‘Mission to Planet Earth,’” Public Administration Review 54, no. 2 (Mar./Apr. 1994): 97–104; W. Henry Lambright, “The Political Construction of Space Satellite Technology,” Science, Technology, and Human Values 19, no. 1 (Winter 1994): 47–69; Richard B. Leshner, “The Evolution of the NASA Earth Observing System: A Case Study in Policy and Project Formulation” (PhD dissertation, George Washington University, 2007); John H. McElroy and Ray A. Williamson, “The Evolution of Earth Science Research from Space: NASA’s Earth Observing System,” in Exploring the Unknown: Selected Documents in the History of the US Civil Space Program: Volume VI: Space and Earth Science, eds. John M. Logsdon, et al. (Washington, DC: NASA, 2004), 441-73; Richard B. Leshner and Thor Hogan, The View From Space: NASA's Evolving Struggle to Understand Our Home Planet (Lawrence, KS: University Press of Kansas, 2019).

210 nebulous ideas that had been germinating in Earth science communities in the decades leading up to the ESSC’s work (chapter one). All the efforts that went into the promotion of ESS help account for why the specific phrase “Earth system” won out as the way to describe the planet as a scientific object, while phrases like “the coupled land-ocean-atmosphere system” fell into disuse. It helps explain why Google’s ngram viewer charts a dramatic spike in use of the phrase

“Earth system” after 1986, precisely when the ESSC’s two reports were published and the group was most active in promoting its ideas.

Again drawing primarily on Francis Bretherton’s ESSC documents from NCAR, chapter four first examines the ways in which both ESSC members and their collaborators recognized that the ESS research program needed to be “sold” to diverse scientific, political, and public constituencies. The chapter then looks at three broad ways in which the ESSC attempted to achieve widespread appeal. First, the ESSC undertook report preparations and review processes that were time consuming and extensive, as the ESSC sought to include as many external experts in the process as possible. It aimed to produce reports that demonstrated scientific rigor but that were also exciting and clearly conveyed a sense of social urgency. Second, the ESSC placed a great strategic emphasis on the importance of aesthetic, rhetorical, and visual considerations for its published materials and presentations. This emphasis is evident in the prominent role played by graphic designer Payson Stevens in developing an Earth system science “brand.” Lastly, the

ESSC employed mechanisms and strategies to spread knowledge about ESS and generate new

(or shore up existing) interest groups. These mechanisms ran the full gamut from the more formal to the informal. They included the published reports, presentations, the development of mailing lists, popular media coverage, a press conference that aired live on CNN, ESS posters, and even ESS t-shirts. Examining all these efforts helps explain the popularity of the “Earth

211

system” as a way of understanding and studying the Earth, even if the ESSC’s specific Earth

system science research program was less enthusiastically received.

THE FIRST STEP

Available ESSC documents leave little doubt that members and affiliated experts were acutely

aware that, to ensure the success of the ESS research program, they would need to convince

many different constituencies that it was worth the time and the resources. Nothing like this—on

such a global scale, with this depth of interdisciplinarity, with this much reliance on Earth

observing satellites and numerical modeling, costing such a large amount of money—had ever

been proposed before for the Earth sciences. The ESSC Working Framework (1984) admitted,

quite frankly, that, “The task before the Committee is daunting, requiring the establishment and

effective presentation of a consensus of many diverse interests, as well as a realistic resolution of

many technical and program issues within a total level of effort that is severely constrained by

budgetary and institutional factors.” This meant that, “The conclusions of the ESSC effort must

also be clearly communicated in a timely and persuasive manner.”6 Early notes on the formation

of the ESSC indicate the importance of “advocacy” and that the committee must “sell, sell, sell”

the ESS research program, taking a page from David Mamet’s Glengarry Glen Ross.7 In a less

frenetic manner, Francis Bretherton emphasized the importance of “selling” the ESS program at

the ESSC’s fifth meeting at JPL in January 1985. It needed to be sold, “To scientific colleagues

6 ESSC Working Framework, Sep 1984, ESSC/Bretherton Committee File, Box 18042, NASA HRC, Washington, DC. 7 Notes, n.d., ESSC/Bretherton Committee File, Box 18042, NASA HRC, Washington, DC. Mamet’s Glengarry Glen Ross made its world premiere at the National Theatre in London in September 1983, and then premiered on Broadway in March 1984. See: Josh Ferri, “Expletives, Awards and Star Power: Why Glengarry Glen Ross Sells as a Modern American Classic,” accessed 24 Apr 2020, https://www.broadway.com/buzz/164979/expletives-awards- and-star-power-why-glengarry-glen-ross-sells-as-a-modern-american-classic/.

212

as well as politicians.”8 To achieve these “sales” objectives, the ESSC needed to make their

reports not only scientifically sound and accessible but also exciting.

Jack Eddy, ever the loyal ESSC report reviewer and frank correspondent, offered telling

comments on a November 1985 draft of the Overview report that demonstrate just how extensive

the ESSC’s efforts were to sell its research program. Eddy described the document as, “an

unusual mixture of TV salesmanship (at the start), powerful science statements (in the middle),

and all-too-terse and telegraphic budget curves and tables (at the end).” Eddy claimed that the

first part of the report, “reads like a revival meeting sermon” or an “advertising brochure,”

written in the “hackneyed style of a NAT’L GEOGRAPHIC article.” He compared some of the

early “salesmanship” phrasing to, “a mayor’s proclamation of Guinea Pig Week.” Lest ESSC

members take his criticisms as indicative of a hostility to the ESS program, Eddy assured them

that, “I hope also that any who read this know that it comes from one of the [ESS] Apostles, an

admirer of the accomplishments of the ESSC, and one who may not fully appreciate the need for

popularizing the study at this time.” Though critical of the Overview’s tone in parts, Eddy

recognized that this was quite probably a calculated response to the need to “sell” the program,

and recommended simplifying the writing in certain parts: “in places it reads like a NASA

engineering-bureaucratese document.” For anyone not familiar with this genre of writing,

Eddy’s comment was decidedly not a compliment. If the goal was to “sell” and present the ESS

material in as accessible a manner as possible, then Eddy recommended omitting the more

complex Earth system wiring diagram (see below), doubting that, “readers of a document that

begins as popularly-written as this does will appreciate the wiring diagram.”9 Eddy’s point was

8 Notes, [Jan 1985], Folder 28, Earth System Sciences Collection, NCAR, Boulder, CO. 9 Earth System Sciences Overview 86, Draft with Comments, 7 Nov 1985, Folder 37, Earth System Sciences Collection, NCAR, Boulder, CO.

213

that, if the ESSC needed to widely publicize their work, then they needed to be more consistent

in their report, and not shift back and forth between a “selling” rhetoric and arid technical

writing.

When the ESSC documents failed to live up to the ESSC’s “sales” objectives, ESSC

members were also quick to note the issue. John Dutton expressed concerns regarding the

Working Framework itself. It might shape many first impressions of the ESSC, and so it was

important to make them good ones. Though not intended to be an interim report, Dutton worried

that it would still be read as such, and so even this document needed to be as exciting and accessible as possible. Dutton described the existing draft Working Framework as a “dangerous document,” emphasizing that, “If this report [the Working Framework] gives the impression that we are just muddling around and don’t have a focus, then our entire effort will be compromised.”

For Dutton, the ESSC needed to be clear about why the “Earth System view is essential” and why this view was, “worth half a billion a year of the taxpayers’ money[.]” This was something

that, in his opinion, the draft Working Framework failed to do. Dutton viewed the Working

Framework as an appallingly dull document that failed to generate excitement or distinguish

itself from previous reports, and indeed it was in part this past “blandness” that led to the need to

form the ESSC in the first place (chapter three). Dutton insisted that the present Working

Framework was, “not really very much to hang a hat on. I would think we would want to go to

the world with something that generates more excitement and is somewhat crisper.” He

recommended having, “a stirring wrap-up, something to salute, before we trundle off to the

details of working groups, members, etc.” As it stood, “The document itself is not nearly ready

214

to go public[.]”10 Some of Dutton’s suggestions were incorporated into the final Working

Framework, though his final opinion has not been preserved. That said, the ESSC’s Working

Framework did not receive widespread (if indeed any) circulation to the general public, which

perhaps indicates that the ESSC recognized a certain failure or lack in the document.

REPORT PREPARATIONS

The ESSC and affiliated researchers recognized the need for the committee to promote its ESS

research program in such a way that it was exciting and accessible to as many different

constituencies as possible. Its first “promotional” strategy was to ensure that its final reports

were as polished and as inclusive as possible. This contrasts with what happened to NASA’s

previous attempt to formulate a large-scale Earth science research program, the Global

Habitability initiative. NASA organized a short workshop of almost 50 elite scientists in June

1982 to study whether and how NASA should carry out this kind of work. But it was a rushed

job, completed in only a few months in order to be presented at the UNISPACE ‘82 conference

in August, and the ensuing report was authored solely by the workshop chair Richard Goody. It

was not a group effort. It did not involve consultation beyond the workshop participants. It did

not provide many details on what a Global Habitability program might concretely look like. As

a result, it received a hostile reception from the international community (chapter two).

Part of what NASA officials learned from that process was to form the ESSC through its

Advisory Council and to give the ESSC adequate time and resources to develop and present its

work. Towards the end of their activities in May 1987, Berrien Moore reminded fellow ESSC

members of the importance of producing a thoroughly polished report and to make sure all small

10 Letter from [John Dutton] to Francis [Bretherton], 26 Aug 1984, Folder 80, Earth System Sciences Collection, NCAR, Boulder, CO.

215 issues were addressed: “precisely because I think the report is so valuable, it must not be flawed by minor items. In fact, it should be so polished that it would reflect the vainest star (the 7th one--visible only with a telescope) in the Pleiades!!”11 Other ESSC members agreed and refused to produce what they considered to be subpar products simply to meet deadlines. Better to delay publication than to send out inferior material. The ESSC recognized the need to produce

“polished” reports that were well-written and compelling, involving not just its internal members, but extensive external editors as well.

Recall the “Words of Wisdom” (WOW) document that collected sayings attributed to

Dixon M. Butler (chapter three). Part of Shelby Tilford’s OSSA Earth Sciences Division, Butler was involved with the conceptual development of both System Z and EOS, and regularly attended ESSC meetings and provided reviewer comments on draft reports. Within both the

EOS and ESSC circles, Butler was known as a character. According to Berrien Moore, “I don’t think Dixon ever slowed down. He was a great enthusiast.”12 This is apparent in Mark Abbott and Seelye Martin’s WOW document, which they wrote to “preserve and maintain the rich, oral tradition of the [EOS Science Steering] committee” as manifested in the 66 unique phrases uttered by Butler throughout their meetings, along with “translations” of these phrases. The

WOW document identified the role of committees in building consensus, particularly for large- scale interdisciplinary programs that will cost “a bundle of money.”13 The humourous point was that, if there are enough committees, then every community member will have served on at least

11 Telemail from B. Moore to F. Bretherton, 6 May 1987, Folder 75, Earth System Sciences Collection, NCAR, Boulder, CO. 12 Berrien Moore, interviewed by Rebecca Wright, Norman, OK, 4 Apr 2011, NASA Johnson Space Center, “Earth System Science at 20 Oral History Project,” accessed 10 Jun 2019, https://historycollection.jsc.nasa.gov/JSCHistoryPortal/history/oral_histories/NASA_HQ/ESS/MooreB/MooreB_4- 4-11.pdf. 13 Mark R. Abbott and Seelye Martin, “Earth Observing System: Stakes in the Ground: An EOS Dictionary,” Oct 1986, Folder 153, Earth System Sciences Collection, NCAR, Boulder, CO.

216 one, and therefore could not object to the ideas being developed. In this way, committees could, for all their tedium, help “sell” a research program.

Butler’s flair for bureaucratic euphemisms was revealed in colourful phrases including gems like:

● “We’re in the right church, moving toward the right pew” ○ translation: “I agree with your argument”

● “Put the meat in the coconut” ○ translation: “Add the substantial issues to the report”

● “The bow wave that I had hoped to break in the fall, will break in the winter” ○ translation: “The report will appear late”

● “By the time this is over, there will be enough wounds to go around” ○ translation: “Writing by committee is hard work”

● “Slipping over the horizon, into never-never land, and off into the sunset” ○ translation: “If we don’t stop fooling around, we’ll never get this thing written”

● “This is moderately mature grist” ○ translation: “If you keep at it, maybe you’ll learn how to write”

● “You will get this bundle on your doorstep” ○ translation: “You’ll receive 10 lbs of reports”

● “They’re singing the right song” ○ translation: “We’ve brainwashed them.”14

All of these phrases, and many more, indicate the hard time put in by committees as they transformed their ideas into reports that could achieve their desired objectives. In the case of the

ESSC, it was to reach and enlist as broad an audience as possible while not sacrificing scientific credibility. That Abbott and Martin took the time to compose this document, and that the majority of Butler’s phrases they preserved revolved around report writing, indicates the

14 Ibid.

217 importance of this activity in committee work and, therefore, in the broader consensus building process.

Internal Reviews

The ESSC fits within the broader culture of post-WWII scientific and political committee work in the US. It was one of countless other similar committees. What makes the ESSC’s work noteworthy is that new technologies enhanced its ability to write and edit collaboratively. This might be taken for granted today, but at the time it was novel and at the cusp of a communications revolution. Personal computers and, especially, electronic mail networks, allowed ESSC members from across the US (and sometimes Europe) to communicate in real time much more easily than previously possible. Electronic mail also allowed members to quickly share drafts of textual material without the delay of regular mail. This had a work- altering effect beyond the real-time exchange of material. It also meant that many different recipients could read a single draft at the same time and provide feedback to all group members.

From its earliest days, the ESSC’s work revolved closely around communications equipment like portable personal computers (precursors of today’s laptops) and telemail services for ESSC members, an approach taken from other groups within NASA.15 Pictures of the

ESSC’s first summer study in June 1984 in Charlottesville, VA, show portable computers in front of some participants, including ESSC chair Francis Bretherton.16 The fact that ESSC members were provided with portable computers at a time when a single, basic machine was being sold for $800 US (around $1,987 US in 2020 dollars) suggests the importance of these

15 Agreement for Electronic Mail, 1 Apr 1984, Folder 130, Earth System Sciences Collection, NCAR, Boulder, CO; Memorandum from Kathy A. Wolfe to NASA Headquarters, 13 Jun 1984, Folder 64, Earth System Sciences Collection, NCAR, Boulder, CO; Memorandum from Stan Ruttenberg to ESSC Committee, Agency Liaisons, Observers, 24 Jan 1984, Folder 129, Earth System Sciences Collection, NCAR, Boulder, CO. 16 Folder 145, Earth System Sciences Collection, NCAR, Boulder, CO.

218

expensive and novel machines for their work.17 Though pricey, these portable computers

promised greater efficiency for the group, and therefore cost-effectiveness in the long-term.

Whether or not these machines actually delivered on this promise is difficult to determine, but

the mere fact that they were procured, at some expense, indicates that at least some believed in

the utility of these machines.

ESSC members were very much aware that they were working in a new way that could

be a template for future work. In the Closer View report, the ESSC devoted a whole chapter to

“Trends in Instrumentation and Technology.” An entire section of that chapter discussed

“Computation Capabilities” that included supercomputing, advanced workstations (meaning

minicomputers and microcomputers, or personal computers), and networking (what we might

call telecommunications). According to the report, “developments are now taking place that

make it possible for increasing numbers of scientists to work together in ways that only a few

years ago would have been unthinkable.”18 Electronic mail’s central purpose served the

committee well: “to permit more efficient use of people’s time. Shortening the time delay

between a request for information and its receipt makes more efficient use of human resources in

the system.”19 The importance of these new technologies emerges from the oral histories of

some of the key ESSC players. Berrien Moore unequivocally asserted that the ESSC’s work

would not have been possible without personal computers and electronic mail. Their work

would have been too episodic and would have relied too greatly on physical meetings when

members were all present, and so the opportunities to collaborate and compromise would have

17 The NEC PC-8200 Personal Computer cost $800, with extra charges for cables, battery pack, RAM, modem, and storage. If one purchased the entire bundle, the cost would be over $1,770, or around $4,397 US in 2020 dollars. See: Brochure, NEC PC-8200 Personal Computer, Folder 129, Earth System Sciences Collection, NCAR, Boulder, CO. 18 ESSC, Closer View, 113. 19 Ibid, 115.

219

shrunken to a completely unworkable level (chapter three).20 Dixon Butler recounted the importance of portable computers and printers even at face-to-face committee meetings, where discussions could be greatly enhanced and expedited if notes of the previous day’s discussions were typed and printed for the following day.21

These new technologies did not always work. Sometimes telemail messages would be

received garbled, or not received at all. The very technologies intended to enhance productivity

and efficiency sometimes had the opposite effect. Personal computers made word processing

and communication easier, and came with different data storage options, but they were also

vulnerable to external forces. A dramatic example of this occurred in September 1986, when

Laura Lee McCauley wrote to the ESSC core writing group regarding a, “very serious

situation[.]”22 Their timeline for completing the Closer View report had fallen behind by an as

yet unknown number of days due to, “the loss of our hard disk” as a result of an electrical storm.

With the telemail subject line “Woe is us,” McCauley reported that, “We were optimistic about

making our first draft deadline for Closer View draft...when we lost our hard disk. We are trying

to recoup but, alas, we must start over on the latest draft.”23 No further details on this incident

are available, but it adds to the explanation of why the Closer View report, which was supposed

to be ready by mid- 1987, was not published until January 1988. Clearly there were tradeoffs to

be made with these new technologies, but the apparent benefits were enough that the ESSC

chose to use them despite the occasions when entire draft reports were lost.

20 Moore interview, 4 Apr 2011. 21 Ibid. 22 Telemail from LL [McCauley] to Payson [Stevens], [9 Sep 1986], Folder 133, Earth System Sciences Collection, NCAR, Boulder, CO. 23 Telemail from L. McCauley to the ESSC Executive Committee, 11 Sep 1986, Folder 133, Earth System Sciences Collection, NCAR, Boulder, CO.

220

The internal review process was often quite intensive, though usually only a smaller

group of ESSC members and support staff were directly involved. This was the core ESSC

writing group (sometimes called the Executive Committee), with Francis Bretherton, Berrien

Moore, John Dutton, Kevin Burke, James Baker, Laura Lee McCauley, and Paul Blanchard as

the primary members. During the internal review process, the ESSC also received aid from

external participants. This was sometimes received informally as ESSC members circulated at

expert gatherings like conferences, workshops, annual meetings, symposiums, and colloquiums.

But external advice was also sought more formally when ESSC members perceived a lack in

their expertise. Notably, the ESSC did not include a hydrologist. According to Geographer John

Wainwright, this was merely a historical contingency since no hydrologist contacted was

available to serve on the committee.24 However, this did not stop ESSC members from reaching

out to the hydrology community for feedback and information. In the summer of 1985 (long

before publication of either the Overview or the Closer View), Robert Ragan, Director of the

Remote Sensing Systems Laboratory at the University of Maryland, met with Berrien Moore and

Kevin Burke of the ESSC while attending a meeting hosted by NASA. Ragan recalled that,

“they felt that although the hydrological cycle was recognized as a key cycle for earth system sciences, this importance was not altogether reflected in the supporting documents.” As a result,

Ragan agreed to prepare a hydrology report for the ESSC, which he forwarded to McCauley for distribution to the ESSC on November 26, 1985. Ragan stated that, “A lot of people in the hydrological science community would sincerely appreciate it if this report could get into the proper channels of ESSC.” Titled “Strategies to Estimate the Global Hydrological Cycle: A

24 John Wainwright, “Earth-System Science,” in A Companion to Environmental Geography, eds. Noel Castree, et al. (Malden, MA: Wiley-Blackwell, 2009), 149.

221

Report to the Earth System Science Committee,” McCauley dutifully complied with the

distribution request.25

Part of the rationale for establishing ESSC Working Groups comprised mainly of non-

ESSC members was to expand the consultation process during the internal report preparation stages. If concerns could be uncovered and dealt with early on, this could drastically reduce the time required for the external review process for the final reports. After the 1984 summer study,

Bretherton wrote to Barry Raleigh asking him to chair the Geophysics Working Group.

According to Bretherton, the “basic functions” of the ESSC’s working groups were two-fold: “to

assist the ESSC with specific questions to which we really need to understand the

answers…[and] to establish a basis of communication between the ESSC and its various

constituencies, to facilitate the achievement of a broad concensus [sic].”26 Throughout the

ESSC’s activities, therefore, there were mechanisms in place by which non-members were able to voice opinions and raise concerns over the ESS research program being developed. Final say over working group membership did rest with Bretherton, but at the very least these working groups provided a means through which external ESSC experts could be involved in the development stage of ESS, something identified as crucial after the Global Habitability experience.27 And it provided the ESSC with another mechanism to build and maintain

community support, as illustrated jokingly by Abbott and Martin’s WOW document.

25 Letter from Robert M. Ragan to Ray Arnold, 7 Aug 1985, Folder 84, Earth System Sciences Collection, NCAR, Boulder, CO. 26 Telemail from F. Bretherton to B. Raleigh, 25 Jun 1984, Folder 86, Earth System Sciences Collection, NCAR, Boulder, CO. 27 Bretherton claimed that this was to give working group chairs, “some protection over issues of balance, etc.” However, given his propensity to micromanage, this could be another manifestation of Bretherton’s desire to remain part of all decision-making processes. See: Telemail from F. Bretherton to B. Raleigh, 25 Jun 1984, Folder 86, Earth System Sciences Collection, NCAR, Boulder, CO.

222

External Reviews

Once the ESSC members reached general agreement about the Overview and Closer View report

drafts, these documents were sent out to external experts for review. This was a crucial part of

the process, so much so that both the ESSC and NASA officials were willing to delay

publication of the reports until these external reviews could take place. When Burt Edelson,

Associate Administrator for NASA’s OSSA, attended the ESSC’s seventh meeting in September

1985, Bretherton informed him that the ESSC could not meet its initial timeline of two years to

produce its report (at this time it was still only one report). McCauley summarized this

discussion, noting, “Edelson accepted the recommendation that more time is needed to flesh out

the report and make sure that the drafts are circulated and given community review[.]”28 That a

NASA official was willing to accept this delay, and therefore commit more resources to the

ESSC, indicates just how much weight the agency placed on the external review process.

Bretherton acknowledged the usefulness of external reviews to Jack Eddy in a letter from 22

August 1988. The process was important not just because it allowed individuals from different

Earth science communities to contribute to the framing of the ESS program, but also because,

“The acquaintances established during this time provide an important network for future

interdisciplinary research activities.”29 This external review process certainly helped prepare the

report but, according to Bretherton, it also helped build a foundation for subsequent

interdisciplinary research.

To facilitate this external review process, McCauley and other support staff devoted

significant amounts of time to the compilation of lists of potential reviewers. Potential reviewers

28 ESSC Update No. 4, [1 Oct 1985], Folder 71, Earth System Sciences Collection, NCAR, Boulder, CO. 29 Letter from Francis P. Bretherton to John Eddy, 22 Aug 1988, Folder 147, Earth System Sciences Collection, NCAR, Boulder, CO.

223

included: ESSC working group members, NASA personnel, NAC members (mostly non-NASA

individuals), members of other relevant federal government agencies (e.g. NOAA, NSF, USGS,

Office of Science and Technology Policy [OSTP]), non-government institutions (e.g

representatives from NCAR and the International Institute for Applied Systems Analysis),

members of the NAS and Space Science Board (SSB), university scientists (e.g. Richard Goody,

Michael McElroy, Lynn Margulis), or simply distinguished experts (e.g. , Michael

Collins).30 All individuals on these lists received the opportunity to comment on the ESSC

report drafts, but not all chose to do so. Bretherton reported that, for the Closer View review, the

ESSC received 41 responses out of a possible 253, which is a response rate of just over 16%.

Bretherton relayed this somewhat disappointing information in a positive manner to the ESSC executive committee, noting that, “The overall tone was favorable, with many constructive suggestions on detail.” Suggested revisions included anything from minor typos and editorial

improvements to more substantive factual corrections and issues relating to the ESSC’s

recommended program.31

Not everyone agreed that the ESSC review process was as rigorous and conscientious as

it should have been. Oceanographer Carl Wunsch, for one, was incensed about the short time

given to reviewers to scrutinize the report. In a terse telemail to McCauley, Wunsch stated that,

“The request for a review on a one week time scale, for a report so long in preparation, is

derisory.”32 Wunsch didn’t even sign his name at the end of his message, perhaps indicative of

30 Folder 75, Earth System Sciences Collection, NCAR, Boulder, CO; Memorandum from F.P. Bretherton to NASA Centers, 7 Jan 1987, Folder 74, Earth System Sciences Collection, NCAR, Boulder, CO; ESSC Reviewers, 17 Apr 1987, Folder 77, Earth System Sciences Collection, NCAR, Boulder, CO; ESSC Reviewers, 1 Jun 1987, Folder 156, Earth System Sciences Collection, NCAR, Boulder, CO. 31 Telemail from F. Bretherton to the ESSC Executive Committee, 11 May 1987, Folder 116, Earth System Sciences Collection, NCAR, Boulder, CO. 32 Telemail from C. Wunsch to L. McCauley, 25 Apr 1985, Folder 48, Earth System Sciences Collection, NCAR, Boulder, CO.

224

his level of frustration. McCauley, at her diplomatic best, attempted to explain why the

turnaround time was so short: “We [the ESSC] realize that we are asking alot, albeit too much.

Yet we have many constraints due to publication deadlines. We had hoped to get it to reviewers

sooner, but it has taken a great effort to get it to this point. Any comments you can give us--

perhaps on just the issues which speak to your expertise--would be appreciated.”33 Wunsch does

not appear to have responded. However, despite this overall problem, many reviewers provided

extensive comments, making in-line comments on the draft ESSC reports with pen and also

providing general feedback comments. The overall response was, as Bretherton reported,

generally positive, with many commending the committee on their achievement.34 No one

reported finding any fault in thinking about and studying the Earth as a system, even if they did

take issue with some of the scientific prioritization (chapter three).

For the ESSC, it was important not just to include as many participants as possible in the

report review processes, but also to publicly note the inclusion of these contributors. If the

ESSC’s recommendations could be shown to be a result of consultations with broader expert

communities, then this would provide them with a stronger justification for their recommended

ESS research program. In a letter to Joan Huffman of Systematics General Corporation

(contracted to NASA for technical administration and office support) on May 13, 1986,

McCauley noted that when Huffman would meet with a representative from Discover Magazine

about the ESSC, Huffman should be sure to emphasize the collaborative nature of the

committee’s work: “the rationale we are presenting and program we are recommending was arrived at by a committee and...it is supported by the community and the relevant [federal]

33 Telemail from L. McCauley to C. Wunsch, 27 Apr 1985, Folder 48, Earth System Sciences Collection, NCAR, Boulder, CO. 34 Folder 48, Earth System Sciences Collection, NCAR, Boulder, CO.

225

agencies--it is bigger than any single individual person or agency.”35 The concluding remarks

for the Overview echo this emphasis on a broad consensus among relevant experts: “The present, two year study of Earth System Science brought together representatives from a wide variety of

Earth-science disciplines and helped to establish new channels of communication and understanding among them. As a result, the approach of Earth System Science is rapidly becoming supported by a broad consensus throughout the Earth-science community.”36 This

collaborative approach helped ensure transparency and scientific accuracy, but it was also a

strategy for building community consensus and political support. The strategy at work, as

evidenced by the ESSC’s approach to its work (and jokes about it) was that if the whole Earth

sciences community had been “brought into” the project by contributing to it in some capacity,

then there would be few, if any, objections to ESS receiving substantial government funding, at

least from the relevant scientific communities.

BRANDING EARTH SYSTEM SCIENCE

Not only did the ESSC invest much effort in preparing the text and content of its reports, but

members recognized the importance of their documents’ visual and aesthetic elements. Many

external scientists, engineers, and politicians who were sent copies of the ESS reports

commented that they were visually engaging. A German meteorologist observed that the reports,

“attract the reader by the appealing layout and the high quality of the illustrations.”37 A Closer

View external reviewer expressed his favourable impression of the document, noting that, “I look

35 Letter from Laura Lee [McCauley] to Joan [Huffman], 13 May 1986, Folder 71, Earth System Sciences Collection, NCAR, Boulder, CO. 36 Earth System Sciences Committee (ESSC), Earth System Science: A Program For Global Change: Overview (Washington, DC: NASA, 1986), 46. 37 Letter from Fritz Kasten to Stan Ruttenberg, 4 Oct 1988, Folder 33, Earth System Sciences Collection, NCAR, Boulder, CO.

226 forward to a copy when the colors are added!”38 Another reviewer requested that he be sent an original, rather than xerox, copy of the Closer View since, “Xerox doesn’t do justice to the pictures and graphs[.]”39

That the ESSC was acutely aware of the importance of aesthetics is supported in a variety of ways. Correspondence on this topic permeates almost all of Bretherton’s ESSC files. When sending the reports to the US Government Printing Office (GPO), they were sent as “a level 1 job,” which was the “highest GPO quality.”40 In a note to Bretherton, Tilford, and the ESSC

Executive Committee in early 1986, McCauley noted that, “It is NOT POSSIBLE to get the

OVERVIEW with its present length out in the sort of classy form needed to grab the congressional audience.” She asked if they should, “go with an ‘interim’ interim brochure, something about 16 pages, including some of the images, colors, icon/symbols that will be in the final, published Overview.”41 Ultimately the answer was “yes,” and the ESSC published the

Preview brochure in early 1986 in order to get something simple yet “classy” ready quickly for external circulation. There were three ways that these aesthetic and visual considerations were made manifest: in the hiring of a graphic design firm (Payson Steven’s InterNetwork), the creation of a distinctive brand for ESS material, and the development of the ESSC’s key visual representation, the wiring diagram of the Earth system.

38 Telemail from T. Spence to L. McCauley, 4 May 1985, Folder 48, Earth System Sciences Collection, NCAR, Boulder, CO. 39 Letter from Gordon McBean to Francis Bretherton, 13 Jan 1987, Folder 149, Earth System Sciences Collection, NCAR, Boulder, CO. 40 Telemail from P. Stevens to L. McCauley, 24 Mar 1986, Folder 153, Earth System Sciences Collection, NCAR, Boulder, CO. 41 Note from Laura [Lee McCauley] to Shelby [Tilford], et al., [1986], Folder 101, Earth System Sciences Collection, NCAR, Boulder, CO.

227

Payson Stevens and InterNetwork

Payson Stevens—along with other ESSC support staff like McCauley and Blanchard—played a

key role in the development of the ESSC reports and other material. McCauley herself lavishes

praise on Stevens (and Blanchard) for transforming the ESSC’s, “work into a ‘product.’”42 Like

other non-scientist members of the ESSC, Stevens’ involvement has not been examined in any of

the sparse historiography on ESSC. This may be due to the paucity of archival material available

on Stevens’ work in NASA’s Historical Reference Collection (the principle repository used by scholars for ESSC materials up to now). But by exploring Stevens’ role as revealed in documents contained in Bretherton’s ESSC files from NCAR, we can better understand the depths of the ESSC’s commitment to producing effective material for general consumption.

Stevens would be the driving force behind many of the “selling” initiatives the ESSC undertook to promote ESS.

With a background in molecular biology and biological oceanography as well as graphic design, Stevens gained extensive experience in preparing media presentations of scientific findings. After completing some graduate work at the Scripps Institution of Oceanography under the guidance of Roger Revelle, Stevens moved permanently into multimedia science

communication. Commenting on Stevens’ subsequent work, Revelle acknowledged that Stevens

was, “a pioneer in this field of combining science and art and I’m enthusiastic about it.”43

Bretherton recognized Stevens’ centrality to the ESSC in a letter to NASA’s Tilford about

Stevens’ contributions to the Overview report: “He made major contributions in clarifying the

basic approach toward communicating with a broad audience. The design of the Overview both

42 Letter from Laura Lee McCauley to Earth System Sciences Committee Members, 31 May 1986, Folder 29, Earth System Sciences Collection, NCAR, Boulder, CO. 43 Newspaper article by Joan Levine, “The View from Out There…” The Tribune, 25 Oct 1983, Folder 101, Earth System Sciences Collection, NCAR, Boulder, CO.

228

in concept and execution were his responsibility. To do this he had to work effectively with a

broad community of scientists. His energy and enthusiasm were infectious...and I appreciated

the opportunity to work with him.”44 Throughout the ESSC’s work, Stevens was described as having a memorable personality, full of enthusiasm and “big picture” ideas. Kathy Wolfe from

SGC called him “a flashy character[.]”45 Nancy Ann Brewster from the NSF wrote to McCauley

that, “I loved meeting Payson [Stevens], what a character, and must be fun to work with.”

McCauley responded that, “yes Payson is a kick to work with[.]”46 Berrien Moore described

Stevens as a, “very creative person…[who] had the ability to pull all of this material together

with very good graphical imagery and it [the ESSC report] became a really monumental

document” because of Stevens’ efforts.47

Stevens was a personable and expressive writer (his telemail correspondence is rife with

capital letters and exclamation points), and an eloquent speaker (so much so that he presented his

ideas on science communication and global warming at a TED talk in 1990).48 By the late 1980s

Stevens’ reputation was such that he was invited to make a presentation at the Sundance

Symposium on Global Climate Change held from August 23 to 26, 1989. In a nod to the

warming of the Cold War brought about by Mikhail Gorbachev, the symposium was nicknamed

“Greenhouse Glasnost” since it brought together over 200 scientists and other interested parties

from the US and USSR to discuss climate change issues and make recommendations, albeit not

very specific ones, to the Soviet and American governments. Funded in part by the actor Robert

Redford’s Institute for Resource Management, Stevens joined a well-recognized group of

44 Letter from Francis Bretherton to Shelby Tilford, 3 Nov 1986, personal correspondence with Payson Stevens. 45 Telemail from K. Wolfe to L. McCauley, 19 Feb 1986, Folder 38, Earth System Sciences Collection, NCAR, Boulder, CO. 46 Telemail from LL [McCauley] to Nancy Ann [Brewster], 27 June 1986, Folder 137, Earth System Sciences Collection, NCAR, Boulder, CO. 47 Moore interview, 4 Apr 2011. 48 “TED2 Talk on Global Warming,” accessed 19 Oct 2019, http://energylandscapes.com/ted2-1990-talk/.

229

participants that included Redford, James Lovelock, NCAR’s Stephen Schneider, former UCAR

president Walter Orr Roberts, James Hansen, Carl Sagan, and Paul Ehrlich.49 After the

conference, Redford wrote to Stevens with words of praise: “the creative energy and imagination

you invested was obvious. It was a tremendous demonstration of putting high technology to its

highest possible use - combining art and computers to educate influential decision-makers about one of the most serious issues facing humanity.”50 In April and May 1990, Senator Al Gore

chaired the Interparliamentary Conference on the Global Environment that brought together

experts and politicians to discuss the many environmental challenges facing the planet. Part of

the proceedings was a keynote address by Carl Sagan and a “hypermedia presentation” by

Stevens.51

Stevens formed InterNetwork Inc. (INI) in 1981 to provide advice and services about

how to convey scientific information to broader, lay audiences using a variety of media,

including film, print, computer graphics, and posters. The company would go on to receive the

Presidential Award for Design Excellence from President Bill Clinton in 1994.52 In the early

1980s, INI received a contract from NASA/JPL to produce a series of images that presented

oceanographic data collected from Earth observing satellites superimposed onto global and

regional images of the planet. This kind of image is commonplace today. Indeed, it is hard to

49 Robert Reinhold, “Summit of Sorts on Global Warming,” New York Times, 27 Aug 1989, accessed 19 Oct 2019, https://www.nytimes.com/1989/08/27/us/summit-of-sorts-on-global-warming.html; “Greenhouse Glasnost Conference, Robert Redford’s Sundance, 1989,” Payson R. Stevens, accessed 19 Oct 2019, https://www.paysonrstevens.com/science-communication/internetwork-media-incinm/redford-sundance- conference/; Terrell J. Minger, ed., Greenhouse Glasnost: The Crisis of Global Warming (New York: Ecco Press, 1990), 291-2. 50 Letter from Robert Redford to Payson R. Stevens, 8 Sep 1989, accessed 19 Oct 2019, https://www.paysonrstevens.com/wp-content/uploads/2018/05/Redford-Ltr-INI-1989.pdf. 51 “The Interparliamentary Conference on the Global Environment,” Payson R. Stevens, accessed 19 Oct 2019, https://www.paysonrstevens.com/wp-content/uploads/2018/07/Gore-Interparliamentary-Conference-on-Global-Env- Agenda-April-1990-HiLite.pdf. 52 “Science Communication,” Payson R. Stevens, accessed 19 Oct 2019, https://www.paysonrstevens.com/science- communication/.

230

imagine how satellite data might be presented except by superimposing them on a map of the

world, using different colours to stand for different data measurements (e.g. carbon dioxide

concentrations, ozone levels, or temperature averages). These are now the familiar images that

populate news pieces on climate change and other environmental crises. But these images did

not simply appear in tandem with the first satellites in 1957. From the beginning, there were

photographs (or television images) captured of the Earth from space that were shown on, say, the

evening local news or in an article. However, early satellite data were presented in tables and

graphs with little attempt to arrange the data in aesthetically compelling or accessible ways.53

This continued into the early 1980s, up to INI’s contract with NASA/JPL.

In 1979, the director of NASA’s Earth Observations Program hired physical oceanographer W. Stanley Wilson to create a comprehensive ocean remote sensing program at the agency in the wake of the successful but short-lived Seasat oceanographic satellite.54 Wilson

recalls that at the time there was a recognition of the need to display satellite data in a way that

would readily reveal patterns that might not be apparent with data simply listed in tables or

displayed in graphs. Wilson hired Stevens’ INI specifically for this reason, to present data from

Seasat and Nimbus-7 satellites that would be intelligible to disparate communities of

practitioners, many of whom did not use satellite data in their research. A new way of

representing satellite data might help build a community of data users and thereby promote

funding for future satellite data projects. Wilson wanted to move away from the presentation of

satellite data in tabular form because it took too much interpretive effort to reveal underlying

53 Personal correspondence. 54 Erik M. Conway, “Drowning in Data: Satellite Oceanography and Information Overload in the Earth Sciences,” Historical Studies in the Physical and Biological Sciences 37, no. 1 (2006): 140.

231

patterns and so was not attention grabbing. For the same reason, he also wanted to move away

from images that required too much explanatory text.55

An example Wilson did not want to imitate comes from the IGY. The IGY was a

worldwide interdisciplinary scientific endeavour to collect coordinated geophysical data on the

Earth’s oceans, lands, and atmosphere from July 1957 to December 1958. Promotional posters

advertising the enterprise were cluttered depictions of various aspects of the planet and its

environs, and the kinds of scientific activities occurring in each area (e.g. Earth, oceans, poles,

space, sun, and weather/climate). The posters included depictions of mythical figures like

Poseidon and inspirational quotations from Romantic poets, the Bible, and Leonardo da Vinci

(Figure 4.1). They also contained smaller images of different features of the scientific

experiments or data collection that would be undertaken, each with a number beside it. A viewer could use this number to look up the image in an accompanying booklet to read a description of what was being depicted.56 Though these were beautiful artistic renderings, the IGY posters

were neither intuitive or straightforward for many contemporary viewers. Hence the additional

interpretive booklet to help decode the representations. By contracting Stevens and INI, NASA

officials like Wilson hoped to create images that would avoid this extra interpretive step. Wilson

wanted to develop images containing satellite data that could be understood at a glance with little

need for further elaboration and certainly no need for an explanatory booklet.57

55 Personal correspondence. 56 ‘Planet Earth’: The Mystery With 100,000 Clues ([Washington, DC]: National Academy of Sciences, 1958); “IGY ‘Planet Earth’ Posters and Booklet,” National Academy of Sciences, accessed 19 Oct 2019, http://www.nasonline.org/about-nas/history/archives/milestones-in-NAS-history/international-geophysical-year- files/igy-picture-galleries/planet-earth.html. 57 Personal correspondence.

232

Figure 4.1. NAS International Geophysical Year poster (1958) depicting scientific activities in the world’s oceans. (Reprinted with permission from the National Academy of Sciences, courtesy of the National Academies Press, Washington, DC)

This was precisely why NASA commissioned INI to produce images of oceanographic satellite data. According to Wilson, all of NASA’s previous outreach documents for satellite

233

data had been simply “sales pieces” that contained no actual observational data from satellites.58

INI’s work in the early 1980s, for the first time, incorporated this data into material for both

scientists and the general public. Whether intentional or not, Stevens’ work on this project

adopted many of the suggestions for “graphical excellence” made by statistician Edward Tufte in

The Visual Display of Quantitative Information (1983).59 For Tufte, visual representations of

data should serve a critical function in the intelligible communication of complex and large data

sets not just to the public, but to specialists as well. Tufte recommends a number of guiding

principles that help achieve this goal: reducing unnecessary ink, removing non-essential

decorative material (“chartjunk”), and merging images, numbers, and words in a way that does

not require much head movement and can convey multiple kinds of information in a single

glance.60 According to Tufte, an excellent graphical display should give, “the viewer the

greatest number of ideas in the shortest time with the least ink in the smallest space.”61 This meant, among other things, that one should, “present many numbers in a small space,” “make large data sets coherent,” “encourage the eye to compare different pieces of data,” and “reveal the data at several levels of detail, from a broad overview to the fine structure.”62

Stevens deployed Tufte’s “graphical excellence” principles into the realm of satellite

data, an area that Tufte did not discuss. The final products for Stevens’ NASA/JPL project were

58 Personal correspondence. 59 Edward Tufte, The Visual Display of Quantitative Information, second edition (Cheshire, CT: Graphics Press, 2001). Whether or not Payson Stevens relied on Tufte’s work, at least one ESSC report reviewer cited Tufte. David Martin from the University of Wisconsin, Madison’s Space Science and Engineering Center explicitly recommended that the ESSC consult Tufte’s book as a resource for developing high-quality graphics. See: Letter from David W. Martin to Laura Lee McCauley, 27 Apr 1987, Folder 48, Earth System Sciences Collection, NCAR, Boulder, CO. 60 One of Tufte’s favourite examples—what he says “may well be the best statistical graphic ever drawn”—is Charles Joseph Minard’s 1869 combination of a data map and time-series to portray the chronological sequence of Napoleon’s troop loss in Russia in 1812. Minard plotted multiple variables (army size, location on a 2-dimensional surface, direction of movement, and temperature) on a single page to great effect. See: Tufte, The Visual Display, 40-1. 61 Tufte, The Visual Display, 51. 62 Ibid, 13.

234

five colour posters with blue/slate backgrounds that depicted oceanographic satellite data for:

Antarctic sea ice, marine winds, phytoplankton abundance, sea surface structure, and sea surface

topography.63 With their ocean-like blue/slate colours, each poster was readily identifiable as

part of the series NASA called “Oceanography From Space.” On each poster, satellite data had

been translated into specific colours that were superimposed upon images of either the entire

planet or a part of the planet. For instance, the poster depicting “phytoplankton abundance”

placed satellite data from Nimbus-7’s Coastal Zone Color Scanner (which measured

phytoplankton chlorophyll concentrations) onto a map of the mid-Atlantic ocean off the eastern

coast of the US (Figure 4.2). Doing so revealed not only the higher concentrations of

phytoplankton on the continental shelf, but also the broad contours of the Gulf Stream as

phytoplankton drift with the ocean currents.64 These patterns popped out of this visual

presentation in a way that they arguably would not have had the data been presented in

traditional tables. Tufte suggests that making patterns in quantitative data visually self-evident is one of the primary purposes of graphically presenting quantitative data. Using examples like

Edmond Halley’s 1686 map showing trade winds and John Snow’s “famous dot map” linking cholera cases to a particular water pump in nineteenth-century London, Tufte argues that graphical displays can “be more precise and revealing” than conventional presentations of statistical data in tables. Differences and patterns can be made manifest with less effort.65

63 A sixth poster illustrated previous and future satellite remote sensors and platforms. 64 “Oceanography From Space,” Payson R. Stevens, accessed 19 Oct 2019, https://www.paysonrstevens.com/science-communication/internetwork-incini/oceanography_from_space/. 65 Tufte, The Visual Display, 23-4.

235

Figure 4.2. NASA’s Oceanography from Space poster for “Phytoplankton Abundance” (1983) as designed by Payson Steven’s InterNetwork graphic design company. (Design by InterNetwork, Inc | Payson R. Stevens; reprinted with permission from NASA)

In an article for San Diego’s Tribune from 1983, Stevens described his approach to

merging science and art in the Oceanography From Space posters: “A new gallery of artistic

ocean imagery is evolving, and scientists are generating its works. They use satellites as their

brushes, computers for canvas, and statistics for paints. The satellites gather the images, the

computers add colors.”66 By combining scientific data with an artist’s aesthetic sensibility,

Stevens could present complex and large data sets in an intelligible and attention-grabbing way to expert and general audiences alike. Stevens’ work on the Oceanography From Space posters proved enormously influential, not just in terms of the presentation of satellite data but also for

66 Newspaper article by Joan Levine, “The View from Out There…” The Tribune, 25 Oct 1983, Folder 101, Earth System Sciences Collection, NCAR, Boulder, CO.

236

his specific work with the ESSC. It provided him with experience in the attractive display of

scientific information and would contribute significantly to his creation of a visual “brand” for

the ESSC material.

Brand Continuity

By hiring Payson Stevens and InterNetwork, ESSC members demonstrated their keen awareness

of the importance of widely distributing and promoting the committee’s materials. As early as

December 1985, Stevens and INI sent the ESSC design and packaging considerations for the

Overview report. After defining the diverse audience for the publication (the public, students,

scientists, and officials in governmental and non-governmental agencies), INI recommended creating what Stevens later referred to as a “brand” for ESS, an aesthetic style that would make all publications and other assorted media material immediately recognizable as part of the ESS program. As sociologist Liz Moor claims in The Rise of Brands (2007), branding is a difficult thing to define since it can vary depending on the contexts in which it takes place and the actors involved.67 In its most abstract, branding, “entails the effort to pattern information—and to embed informational qualities in material goods—in order to organize experiences and perception in line with particular strategic ends.”68

Branding has a long history (for instance the branding of livestock traces back at least to

Ancient Greece), but in the nineteenth century, companies increasingly began to rely on branding

as a way to differentiate their products from competitors and guarantee quality.69 By the 1980s, branding expanded to include all the ways that a corporate entity presented itself to the public to

67 Liz Moor, The Rise of Brands (New York: Berg: 2007). 68 Ibid, 9. 69 Ibid, ch. 2.

237 promote not just physical products, but more intangible values, meanings, feelings, and lifestyles. For instance, Apple products took on new connotations after the 1984 airing of its

“Big Brother” advertisement for the new Macintosh personal computer.70 Directed by Ridley

Scott, the ad is set in a dystopian future of conformity where mindless citizens unquestioningly submit to an image of an authoritarian figure on a screen. The scene is disrupted by a lone female runner with a sledgehammer who destroys the screen and (presumably) liberates the minds of the citizenry. The set of values Apple was attempting to cultivate is unambiguous. To buy an Apple personal computer in 1984 meant more than simply becoming a computer owner.

It meant aligning with a brand that promoted the values of personal autonomy, creativity, and individuality, and by implication not the values of its chief competitor, IBM.

A number of factors laid the foundations for this greater emphasis on branding in the

1980s particularly in North America, including privatization, deregulation, intensifying competition, increasing merger and acquisition activity, deindustrialization (at least in the West), and the rise of a service economy.71 For this dissertation, it is important to note that in the 1980s there was a broad move towards constructing and promoting corporate brands, and away from simply manufacturing specific products. A brand incorporated a company’s products, but it was not reducible to them. According to Moor, “Branding has developed from initially marking property and ownership, and identifying the origin and content of goods, to connoting different types of values, meanings and reputations.”72

Stevens’ emphasis on “branding” ESS is much in keeping with the spirit of the 1980s, a time when corporate branding took on a much greater degree of importance and spread into

70 “Apple 1984 Super Bowl Commercial Introducing Macintosh Computer,” accessed 14 Apr 2020, https://www.youtube.com/watch?v=2zfqw8nhUwA. 71 Moor, The Rise of Brands, 30-5. 72 Ibid, 15.

238 other, non-commercial sectors like charities, environmental groups, and even political parties.73

To be clear, linking Stevens’ ESS branding to a larger 1980s branding zeitgeist is not to impute any nefarious motivations to his work or to delegitimize the project. While Naomi Klein critiques the branding activities of multinational corporations like Nike, Disney, and Starbucks, what she objects to is not “branding” per se, but how it has been used to generate appeal and profits while hiding dubious practices like labour exploitation and environmental degradation.74

Branding in its most basic and abstract sense is not so nefarious. It simply means patterning information to affect experiences and perceptions in order to achieve a particular goal.

This is not fundamentally different from what humans do when they communicate with each other, using language to “control” the thoughts and ideas of others. (For instance, when I say

“chocolate tastes good” to someone, I want my interlocutor to think about chocolate tasting good, hence I want to control their thoughts, even if only in a mundane way.) An act of attempted control is arguably essential for communication.75 Branding is a larger and more coordinated effort to communicate and influence action using a variety of available media, be it for Nike, Disney, Apple, New Labour, breast cancer, Greenpeace, or UNICEF. Branding is neither good nor bad; nor is it neutral.76 The relevant questions to ask about branding are: what are the methods employed and what is the ultimate goal? Answers to these questions enables a case-by-case assessment of branding techniques without impugning all branding by a logically

73 Moor uses the example of Britain’s Labour Party rebranding itself in the 1990s into New Labour as a prominent recent example of non-commercial branding. See: Moor, The Rise of Brands, 4. 74 Naomi Klein, No Logo: No Space, No Choice, No Jobs: Taking Aim at the Brand Bullies (London: Flamingo, 2000). For an example, one need only look at the timing of Nike’s most recent ad campaign involving Colin Kaepernick. It came out on September 5, 2018, while just the previous month four women brought a gender discrimination lawsuit against the company. See: Edward Helmore, “Nike Hit With Lawsuit From Four Women Who Allege Gender Discrimination,” The Guardian, 10 Aug 2018, accessed 19 Oct 2019, https://www.theguardian.com/business/2018/aug/10/nike-lawsuit-women-gender-discrimination. 75 See: Norbert Wiener, The Human Use of Human Beings: Cybernetics and Society (New York: Doubleday, 1954). 76 This is a modification of Melvin Kranzberg’s “First Law” of technology: “Technology is neither good nor bad; nor is it neutral.” See: Melvin Kranzberg, “Technology and History: ‘Kranzberg's Laws,’” Technology and Culture 27, no. 3 (Jul. 1986): 545.

239 suspect “guilt by association” argument. For example, Nike and UNICEF might both use branding techniques but we need not view their methods and goals as comparable or equally in need of critique (or praise). The same can be said for Earth system science as branded by Payson

Stevens in the 1980s.

Ideas for studying the Earth in more interconnected and interdisciplinary ways were nascent but growing from the 1960s into the 1980s amongst Earth scientists (chapter one).

However, apart from the short-lived and unsuccessful Global Habitability initiative, no group had proposed such a large-scale and integrated research program as ESS. With the problematic

Global Habitability experience never far from their minds, ESSC members recognized that it would require substantial work to realize the ESS research program. The ultimate goal was to promote ESS by presenting it as an enterprise invested with certain values and meanings. ESS as a brand promoted a “lifestyle” that incorporated environmental concerns, interdisciplinary Earth science research, Earth observing satellites, computer modeling, interagency cooperation, and international collaboration. If one identified with this scientific lifestyle (whether directly or in principle), then one would, it was hoped, support an ESS research program. Stevens was hired to develop techniques to promote the ESS “lifestyle” in US scientific and science policy communities.

To create an ESS brand, Stevens recommended designing all ESS material using a unified theme that incorporated colour, layout design, and image choices. This was what

Stevens had done with the Oceanography From Space posters and their blue/slate backgrounds.

He went even further now, recommending the development of an ESS “logo” that could accompany (and would therefore identify) all related ESS material (Figure 4.3). He further recommended the development of symbols for the different components of the Earth system (e.g.

240 solid, fluid, biological Earth).77 While this suggestion was never taken up, Stevens’ ESS logo became a reality. It can be found on all of the ESSC’s published materials, and even on ESS t- shirts that Stevens had printed and proposed selling at ESS conference displays.78

Figure 4.3. The ESSC’s Earth system science logo found on its reports and other media products. (Design by InterNetwork, Inc | Payson R. Stevens; reprinted with permission from NASA)

The most obvious bit of branding that the ESSC material contains is its distinctive blue colour. This blue was Pantone Matching System (PMS) 301 Blue.79 It permeates all of its

77 Preliminary Ideas, Design/Packaging Considerations for Overview 86 by InterNetwork Inc., 11 Dec 1985, Folder 101, Earth System Sciences Collection, NCAR, Boulder, CO. 78 Note from Laura [Lee McCauley] to Sarah [Hawkins], 16 May 1986, Folder 30, Earth System Sciences Collection, NCAR, Boulder, CO. Stevens can be seen wearing an ESS t-shirt in a picture taken of him and Robert Redford at the Greenhouse Glasnost conference. See: “Greenhouse Glasnost Conference, Robert Redford’s Sundance, 1989,” Payson R. Stevens, accessed 19 Oct 2019, https://www.paysonrstevens.com/science- communication/internetwork-media-incinm/redford-sundance-conference/.

241

material. It is on the covers of both the Overview and Closer View reports, as well as the

Preview brochure (Figure 4.4). It adorns the folders in which the reports and other material were

packaged for the ESSC press conference and later for mass mailings of materials. It was on the

ESS posters and presentation slides. The ESSC’s letterhead used this blue, and it even served as

the background colour for Stevens’ ESS t-shirts. Time and again, individuals commented on this distinctive feature of the ESSC materials, referring to it variously as the “trademark blue” or “our blue” or “ESSC blue” or “ESS blue.”80 Stalwart ESSC reviewer Jack Eddy referred to ESSC

publications as the “blue reports,” suggesting that this blue had successfully become distinctive

beyond the strict confines of ESSC membership.81 Stevens viewed the maintenance of visual and thematic unity across different media outputs as an important way to link different materials and create a unified project readily recognizable to a number of different audiences. It was a way for the viewer, at a glance, to identify the material and link it to previously-viewed objects

(or begin this process if this was the first viewing). A glance at, say, an ESSC report cover could give the reader immediate insight into the general contents of the document, once the ESS brand

79 Memorandum from Payson Stevens to Shelby Tilford, et al., 27 Mar 1986, Folder 32, Earth System Sciences Collection, NCAR, Boulder, CO; “PMS 301 C,” PMS Color Guide, accessed 19 Oct 2019, https://www.pmscolorguide.com/coated/pms-301-c. 80 Earth System Science Media Support: Preliminary Budgets, [17 Oct 1985], Folder 120, Earth System Sciences Collection, NCAR, Boulder, CO; Memorandum from Payson Stevens to Shelby Tilford, et al., 27 Mar 1986, Folder 32, Earth System Sciences Collection, NCAR, Boulder, CO; Memorandum and Attachments from Laura Lee McCauley to ESSC Members, et al., 15 May 1986, Folder 71, Earth System Sciences Collection, NCAR, Boulder, CO; Memorandum from Payson Stevens to Shelby Tilford, et al., 27 Mar 1986, Folder 41, Earth System Sciences Collection, NCAR, Boulder, CO; Memorandum from Paul Blanchard to Francis [Bretherton] et al., 18 Apr 1986, Folder 118, Earth System Sciences Collection, NCAR, Boulder, CO; Memorandum from Payson Stevens to Paul Blanchard and Laura Lee McCauley, 29 May 1986, Folder 72, Earth System Sciences Collection, NCAR, Boulder, CO; Letter from Francis P. Bretherton to ESSC Members, 6 Apr 1987, Folder 79, Earth System Sciences Collection, NCAR, Boulder, CO. 81 Telemail from J. Eddy to S. Tilford, et al., 14 Mar 1987, Folder 77, Earth System Sciences Collection, NCAR, Boulder, CO.

242 became well-enough known. The ESS blue further linked the material to the coherent set of ideas and values that made up the ESS brand.

Figure 4.4. The “ESS Blue” serves as the background colour for all ESSC material, including its Closer View (1988) report. (Design by InterNetwork, Inc | Payson R. Stevens; reprinted with permission from NASA)

243

For Stevens, his different NASA projects were not self contained but part of a broader

project that involved large-scale Earth science studies of different components of the planet

using satellites. Stevens intentionally developed the ESS theme as a continuation of the

Oceanography From Space colour scheme (even if it was less slate and more blue under ESS).

The deliberate link was both conceptual and concrete. In a telemail message from early January

1986, Stevens outlined some of his thoughts about a promotional strategy for ESS. The last

section of this chapter will look at this strategy in more detail, but here it is important to note

Stevens’ vision for how all of the ESS material should be aesthetically linked together.

According to Stevens, all of this material (brochures, reports, posters, folders, postcards, and

convention displays), “should be married to each other through key design elements (color for

program; logo; typeface; etc.) much the way the ‘Oceanography From Space’ products were.”

He even recommended a series of posters, “analogous to the ‘Oceanography From Space’

series.”82 The glossy images included as part of the Closer View report folder released in

January 1988 depict satellite data (e.g. sea surface topography, Antarctic ice, air-sea interactions, stratospheric ozone) superimposed on maps of part or all of the Earth’s surface. McCauley wrote to Stevens on June 25, 1987 regarding these images and compared them to, “the oceanography from space stuff.”83

Beyond these conceptual links, Stevens envisioned a concrete continuity between ESS and Oceanography From Space material. In preparation for the ESS press conference on June

26, 1986 (see below), Stevens recommended that NASA Headquarters’ Public Affairs Office

82 Telemail from P. Stevens to S. Tilford, et al., 3 Jan 1986, Folder 55, Earth System Sciences Collection, NCAR, Boulder, CO. 83 Telemail from L. McCauley to P. Stevens, 25 Jun 1987, Folder 156, Earth System Sciences Collection, NCAR, Boulder, CO.

244

mount both ESS and Oceanography From Space posters onto foam boards for display.84 Others

involved with the ESSC’s work similarly understood and voiced the importance using aesthetic

continuity between the Oceanography From Space posters and ESS material. Shelby Tilford

defended an extension of Stevens’ contract to work with the ESSC specifically because of the

potential for cohesion between ESSC outputs and previous work done by Stevens’ INI on the

oceanography posters: “it is desirable to maintain continuity in these publications to aid public

association of these efforts--clearly the unique and successful style of design used by INI is not

capable of duplication elsewhere.”85

The Oceanography from Space images had enjoyed a warm reception, and arguably the

ESSC was attempting to leverage this success by co-branding. By linking itself to successful and attention-grabbing imagery, the promise of ESS research might be more easily imagined. All of this demonstrates the concern that went into developing ESS products as things immediately recognizable as part of a larger, overarching brand. The linking of satellite remote sensing, computer modeling, interdisciplinary Earth science research, environmental concerns, and an interconnected planet was not just a research project but a lifestyle choice.

The Wiring Diagram

One of the ESSC’s most important visual and aesthetic activities was the development of the

Earth system wiring diagrams, the conceptual depictions of the structure of the Earth as a system

with interacting components. The most famous of these would eventually be referred to as the

“Bretherton diagram.” Since the publication of Martin Rudwick’s influential 1976 paper “The

84 Telemail from P. Stevens to L. McCauley et al., 12 Jun 1986, Folder 50, Earth System Sciences Collection, NCAR, Boulder, CO; Art List, [26 Jun 1986], Folder 50, Earth System Sciences Collection, NCAR, Boulder, CO. 85 Letter from S.G. Tilford to D.G. Rea, 17 Oct 1986, Folder 77, Earth System Sciences Collection, NCAR, Boulder, CO.

245

Emergence of a Visual Language for Geological Science, 1760-1840”—what Evan Hepler-Smith

calls the “founding document”—visualization and representational practices have become an

important and popular object of study in the history, philosophy, and sociology of science.86

According to Michael Lynch and Steve Woolgar in Representation in Scientific Practice (1988), these representations can take on a variety of forms (including graphs, charts, diagrams, photographs, instrumental data, or computer programs) and can serve a variety of functions (e.g. resemblance, symbolic reference, similitude, abstraction, exemplification, expression). They argue that what links scholarly studies of representations in the sciences is the focus on how representations are used by scientists in particular contexts to support their work.87

Kathryn Henderson and Bruno Latour have examined how representations can be used as

means of persuasion in the building and maintaining of scientific consensus. Henderson’s

“conscription devices” and Latour’s “immutable mobiles” help understand how representations

organize individuals around common goals, and how they present absent or abstract ideas in

ways that are easily readable, understandable, and reproducible. A representation allows for

concepts, techniques, and practices to travel beyond their originating sites.88 While Henderson

and Latour emphasize the unchanging aspect of a representation that draws things and people

together, David Kaiser downplays the significance of immutability in favour of interpretive

86 Evan Hepler-Smith, “Grasping the Technical Image,” Historical Studies in the Natural Sciences 47, no. 1 (2017): 118; Martin J.S. Rudwick, “The Emergence of a Visual Language for Geological Science, 1760-1840,” History of Science 14, no. 3 (1976): 149-95. 87 Michael Lynch and Steve Woolgar, “Introduction: Sociological Orientations to Representational Practice in Science,” in Representation in Scientific Practice, eds. Michael Lynch, Steve Woolgar (Cambridge, MA: MIT Press, 1990), 14. 88 Kathryn Henderson, On Line and On Paper: Visual Representations, Visual Culture, and Computer Graphics in Design Engineering (Cambridge, MA: MIT Press, 1999), 53; Bruno Latour, “Drawing Things Together,” in Representation in Scientific Practice, eds. Michael Lynch, Steve Woolgar (Cambridge, MA: MIT Press, 1990), 50.

246

“plasticity,” at least in the specific context of the dispersion of Feynman diagrams among post-

WWII physics communities.89

Susan Leigh Star’s “boundary object” incorporates immutability and plasticity features.

A boundary object facilitates communication amongst diverse communities through its

multilayered meaning. It is an object or idea that is pliable enough to be used in a variety of

local contexts, and yet robust enough that it maintains a common meaning across different

sites.90 This semantic multiplicity applies to the Earth system concept, making the boundary

object a useful analytical tool to explain how a Marxian “semantic void” might be filled.

Understanding the Earth system as a boundary object helps explain how it traveled beyond the

confines of the ESSC even while the ESS research program encountered more resistance

(chapter five).

The Bretherton diagram served as both the negotiated understanding of the interacting

components of the Earth system and as a consensus building tool. It would also turn out to be

the most memorable and significant product developed by the ESSC. By 2009, when NASA’s

History Office conducted interviews with some scientists and NASA officials associated with

ESS, many had forgotten or misremembered key events and details. Despite this, the one thing

that almost all actors reported was the significance they placed in the Bretherton diagram.91 One

individual—solar-terrestrial physicist and ESSC member Lennard Fisk—went so far as to assert,

somewhat bizarrely, that the ESSC had produced no reports, at least none he had ever been able

89 David Kaiser, Drawing Theories Apart: The Dispersion of Feynman Diagrams in Postwar Physics (Chicago: University of Chicago Press, 2005), 17, 173, 281-2. 90 Susan Leigh Star and James R. Griesemer, “Institutional Ecology, ‘Translations’ and Boundary Objects: Amateurs and Professionals in Berkeley's Museum of Vertebrate Zoology,” Social Studies of Science 19, no. 3 (Aug. 1989): 393. 91 See: Dixon M. Butler, interviewed by Rebecca Wright, Washington, DC, 3 Jun 2010, NASA Johnson Space Center “Oral History Project,” accessed 10 Oct 2019, https://historycollection.jsc.nasa.gov/JSCHistoryPortal/history/oral_histories/NASA_HQ/ESS/ButlerDM/ButlerDM _6-3-10.pdf; Moore interview, 4 Apr 2011.

247

to locate, “But the Bretherton Wiring Diagram has lived forever.”92 In his Earth System Science:

A Very Short Introduction (2016), self-described Earth system scientist Tim Lenton proclaims

the Bretherton diagram to be the “most lasting legacy” of the ESSC. According to Lenton, the

Bretherton diagram provided the “social glue” for interdisciplinary studies of the Earth by

putting, “a whole range of existing scientific subjects—and their associated scientific

communities—together on the same map.”93 Chapter three examined the activities of the

ESSC’s ESMWG and members’ efforts to define what was in and what was out of the Earth

system in order to develop a structural model of the Earth system in the form of a wiring

diagram. That discussion of the diagram was content-oriented. However, ESMWG and ESSC members also took into account aesthetic details that contributed to consensus-building and shaped the final form of the Bretherton diagram.

To refer to the “Bretherton diagram” in the singular is misleading since, in fact, there were two major wiring diagrams that the ESSC included in its final report. The first diagram was a conceptual model of Earth system processes operating on timescales of decades to centuries, while the second was for Earth system processes operating on timescales of thousands to millions of years (chapter three). To complicate matters further, there were actually two versions of the wiring diagram for shorter term processes, a simplified version (Figure 4.5) and a more complex one (Figure 3.1). In its reports, the ESSC clearly viewed the more complex wiring diagram as the scientific centerpiece of its work. That diagram depicted—for processes occurring on timescales of decades to centuries—major components of the Earth’s systems (the

92 Lennard Fisk, interview by Rebecca Wright, Ann Arbor, MI, 8 Sep 2010, NASA Johnson Space Center, “Oral History Project,” accessed 12 Aug 2019, https://historycollection.jsc.nasa.gov/JSCHistoryPortal/history/oral_histories/NASA_HQ/Administrators/FiskLA/Fis kLA_9-8-10.pdf. 93 Tim Lenton, Earth System Science: A Very Short Introduction (New York: Oxford University Press, 2016), 14. Interestingly, Lenton only identifies one ESSC report, the Overview.

248 physical climate and biogeochemical cycles systems) and showed the dominant processes that connected components.

But it was a very detailed diagram to absorb in a single glance, with tiny text and symbols that could overwhelm a viewer. Even a practicing geologist, W.G. Ernst, reported feeling, “non- plussed by all the block diagrams which look like schematic wiring diagrams. Most of them are

(at least to me) not intuitively obvious.”94 Recognizing that not everyone might be familiar with this kind of diagram, the ESSC was careful to explain how the symbols should be interpreted

(e.g. the arrows represented information pathways connecting subsystems, the ovals represented either inputs to or outputs from the Earth system, and the rectangle boxes represent subsystem components).95 Berrien Moore emphasized the need to, “‘walk’ the reader through the figure.”96

Yet even with these viewing instructions, the complex diagram might alienate readers, particularly non-Earth scientists who might include science policy makers. Commenting on a

November 1985 Overview draft, Jack Eddy raised this concern: “I question whether readers of a document that begins as popularly-written as this does will appreciate the [complex] wiring diagram. I would omit it.”97

94 Letter from W.G. Ernst to Francis P. Bretherton, 4 May 1987, Folder 47, Earth System Sciences Collection, NCAR, Boulder, CO. 95 ESSC, Closer View, 28, 30, 74, 95. 96 Telemail from B. Moore to the ESSC Executive Committee, 16 Apr 1987, Folder 136, Earth System Sciences Collection, NCAR, Boulder, CO. 97 Earth System Sciences Overview 86, Draft with Comments, 7 Nov 1985, Folder 37, Earth System Sciences Collection, NCAR, Boulder, CO.

249

Figure 4.5. The simplified Earth system wiring diagram, for timescales of decades to centuries, from the ESSC’s Overview (1986) report. (Design by InterNetwork, Inc | Payson R. Stevens; reprinted with permission from NASA)

The ESSC did not omit the more complex diagram from the final report, but presented it in tandem with a simplified version that could more readily serve as a consensus building device for viewers, expert or otherwise. What eventually gets referred to as the “Bretherton diagram” and becomes the most notable ESSC product is not the more detailed (and arguably more scientifically accurate) wiring diagram but the simplified version of the Earth system that depicts its main components but contains little scientific information on material and energy flows.98

What was most useful and compelling for external readers was not a diagram that had scientific fidelity but a diagram that offered a visual encapsulation of what an integrated Earth system

98 For a recent example, see: Sarah Dry, Waters of the World: The Story of the Scientists Who Unraveled the Mysteries of Our Oceans, Atmosphere, and Ice Sheets and Made the Planet Whole (Chicago: The University of Chicago Press, 2019), ch. 7. Lenton also uses the “Bretherton diagram” to refer to the simplified diagram. See: Lenton, Earth System Science, 14.

250

looked like. The simplified diagram, which was the only wiring diagram to appear in the

ESSC’s first publication (the Overview) in 1986, was the first picture of the Earth system

concept.99 Scientific realism might be useful for certain experts, but it was not necessary or even

desirable in order for the wiring diagram to serve as a consensus-building device for ESS.

One of the ESSC’s objectives in developing the Earth system wiring diagrams was to design a visual image that could be easily understood, one that could draw together Earth scientists by showing how different disciplines fit into ESS. A prominent example from the history of science of using visual representations as a mechanism for achieving a scientific consensus comes from Samuel Edgerton’s analysis of Galileo’s depictions of the moon’s surface in Sidereus nuncius (The Starry Messenger, 1610). Edgerton argues that, rather than creating an

accurate map of the moon’s surface, Galileo instructed his engraver to exaggerate certain

topographical features on the moon to convince readers that the moon’s surface was like the

Earth’s, with mountains, valleys, and seas. Galileo’s images were intended, “as a kind of

kinesthetic expression rather than as a cartographic fact” in support of a Copernican view of the

universe that made the Earth one of the planets and eliminated the perfect superlunary realm in

which there could be no blemishes on the moon.100 As Janet Vertesi argues, Galileo wanted

readers to see the moon as a landscape commensurate to landscapes on Earth so he drew the

moon as this kind of object, with this theory-laden assumption guiding his drawing (what Vertesi

calls “drawing as”).101 Scientific realism was less important than instilling a non-Aristotelian

view of the world.

99 ESSC, Overview, 19. 100 Samuel Y. Edgerton Jr., “Galileo, Florentine ‘Disegno,’ and the ‘Strange Spottednesse’ of the Moon,” Art Journal 44, no. 3 (1984): 228-9. 101 Janet Vertesi, “Drawing as: Distinctions and Disambiguation in Digital Images of Mars,” in Representation in Scientific Practice Revisited, eds. Catelijne Coopmans, et al. (Cambridge, MA: MIT Press, 2014), 17-21.

251

In much the same manner, the Bretherton diagram did not rely on a realistic or

isomorphic mapping between the wiring diagram and the world. What was crucial was to depict

the Earth as an interconnected system containing subsystems in a way that Earth scientists would

find compelling. On the simplified diagram, broad disciplinary areas like “marine biochemistry”

and “ocean dynamics” serve as labels for specific subsystems. This allowed the diagram to serve

as a consensus-building device for ESS by grafting disciplinary consensus onto Earth

subsystems. Scientists use visual representations—be it Galileo’s lunar landscape, Vertesi’s

Mars Rover images, or the Bretherton diagrams—to build communities of like-minded

individuals that coalesce around a particular set of beliefs and practices. Scientific

representations like the Bretherton diagram are often what Hepler-Smith calls “social

achievements,” the end product of group organization and agreement.102 But these representations can also do “social work” by creating new communities or enlisting existing ones.103

That this was one of the functions of the Bretherton diagram becomes apparent when

examining the history of its development. The majority of the work on both the simplified and

more complex versions of the conceptual model of the Earth system for processes occurring on

timescales of decades to centuries was completed at the ESMWG’s second meeting in February

1985. Apart from Burke creating a separate wiring diagram for longer term processes that

appeared in the Closer View (1988), the details contained in the shorter term process diagrams

changed very little after February 1985. One minor change was made to the simplified diagram

to address John Dutton’s concern that an early version of the box representing “human activities”

102 Hepler-Smith, “Grasping the Technological Image,” 123. 103 Janet Vertesi, Seeing Like a Rover: How Robots, Teams, and Images Craft Knowledge of Mars (Chicago: University of Chicago Press, 2015), 244.

252

looked too much like a “WC” figure.104 Subsequently, more realistic human images were added

to this box. The only major revision the ESSC sanctioned for the complex wiring diagram was

to make it visually more compelling. In 1987 Payson Stevens redesigned the complex diagram

to ensure that its type and graphics could, “be easily read when projected as [a] 33mm slide.”105

These were the only revisions despite some external reviewers raising concerns over

some of the diagrams’ details. For instance, cryosphere specialist and ESSC member Wilford

Weeks suggested that one of the external components of the complex wiring diagram contained

in the September 1985 draft of the ESSC report should be altered from “ice cores” to “ice sheets

and shelves.”106 There was plenty of time to incorporate the suggestion into the final report, and

yet it was ignored. This suggests different priorities than scientific realism. “Ice cores” was

shorter and easier to read than the kludgier “ice sheets and shelves.” In another example,

atmospheric scientist and Closer View reviewer David Martin looked at the wiring diagram for

long-term processes and questioned the connections made between the generation and

degradation of topographical relief as well as between sea topography and physical climate

processes.107 There is no record of these questions being answered and the connections

remained in the Closer View’s final version of the diagram. Atmospheric scientist Michel

Verstraete generated a page-long list of scientific problems with a version of the complex wiring

diagram. His criticisms included: a lack of matter-energy conservation for incoming solar

104 Telemail from J. Dutton to ESSC Executive Committee, 9 Jan 1986, Folder 34, Earth System Sciences Collection, NCAR, Boulder, CO. 105 Earth System Science Media Support: Preliminary Budgets, [17 Oct 1985], Folder 120, Earth System Sciences Collection, NCAR, Boulder, CO. 106 [Overview] Editing by W.F. Weeks, [September 1985], Folder 10, Earth System Sciences Collection, NCAR, Boulder, CO. 107 ESSC Closer View Comments from David Martin, [27 Apr 1987], Folder 48, Earth System Sciences Collection, NCAR, Boulder, CO.

253

radiation, the water cycle, and carbon cycle;108 the failure to include thermal pollution among the

human activities that affected the Earth system; and the absence of soil erosion.109 Again, none

of these concerns were addressed. Clearly what was important in the wiring diagrams was not so

much ensuring that every scientific detail was as rigorously recorded as possible, but creating a

visual representation of the Earth system that effectively conveyed the Earth system concept and

could thereby serve to unite Earth science practitioners. Thus, the main goal was not to create an

accurate scientific model of the world. Rather, the aim was to create a visual representation for

heuristic purposes that showed Earth scientists precisely where their research fit into ESS.

The Bretherton diagram served as a heuristic tool for generating consensus about the

Earth as a system and providing a background framework for interdisciplinary cooperation. In a

20-year post hoc oral history, Shelby Tilford described the Bretherton diagram as a “chart” that,

“couples almost everything, every element of the Earth system, going from the sun to the center

of the Earth. That includes the solar input, the effect of the upper atmosphere, lower atmosphere,

the troposphere, et cetera. It covers weather, it covers oceans, it covers land, it covers the solid

Earth. We tried to put together something that would integrate the Earth as an integrated science

program. That’s what the chart does.”110 Tilford’s distant recollection of what components were actually included in the Bretherton diagram is not entirely accurate or clear. For instance, the solid Earth components are not actually part of the Earth system in the Bretherton diagram—

which represents global change on timescales of decades to centuries—but are only inputs to the

system. However, those components were still present in the image, even if not technically part

108 The ESSC was aware that their complex Earth system wiring diagram for processes occurring on time scales of decades to centuries would not maintain balances for matter and energy since some of those fluxes occurred on much longer time scales. See: Earth System Modeling Working Group Meeting Report, Francis Bretherton, 6 Mar 1985, Folder 91, Earth System Sciences Collection, NCAR, Boulder, CO. 109 Note from Michel M. Verstraete, n.d., Folder 6, Earth System Sciences Collection, NCAR, Boulder, CO. 110 Tilford interview, 23 June 2009.

254

of the Earth system on this timescale. The image could, thus, still provide an important integrating role for Earth science practitioners.

Bretherton recounted something similar in personal correspondence with historian Eric

Goldstein. He described the diagram as a “high level systems architecture of a computer model”

for the entire Earth system. Once developed, the diagram could be used to demonstrate how the

different disciplines fit together in the larger framework of studying the Earth as a system:

[I]t became useful because from there on every time I was talking about this program to any group of scientists...the message was this is where you fit in! Your community, your effort—you can only help with this aspect of the measurements, or this aspect of your modeling, which enabled us to infer some things which we can’t measure but nevertheless information that’s needed to say how the system as a whole will look rather than just the fragmented pieces of it.111

In short, this abstract diagram could act as a shorthand for how disciplinary research on the Earth

system could fit into a comprehensive and ultimately integrated framework. Bretherton could,

say, point to the subsystem box labeled “ocean dynamics” to show a physical oceanographer

where his/her work fit into the larger interconnected system, or point an ecologist to the

“terrestrial ecosystems” box, or an atmospheric chemist to the “tropospheric chemistry” box, and

so forth.

For the ESSC, the main utility of the Bretherton diagram was its consensus-building

potential. It linked together the various Earth science disciplines and demonstrated the

interconnectedness of their different research areas. It was an integrator, taking in disparate

disciplinary pieces and returning a large-scale, cohesive, interdisciplinary ESS research program.

Experts in specific areas were still needed, but so too were those who could work across the traditional divisions since the components of the Earth system were not actually isolated from

111 Quoted in: Goldstein, “NASA’s Earth Science Program,” 134.

255

one another. It was previously a practical necessity to work in a compartmentalized fashion, but

with new and improving technologies (like satellites and computers), practitioners could now

more readily engage in interdisciplinary cooperation. Historian and philosopher of science

Chunglin Kwa argues that the ESSC’s work became “especially famous” because of this

diagram: “It was useful because the diagram implicitly identified the sciences that were to be

included in a global research programme and assigned a meaningful role to each of them.”112

Lennard Fisk, somewhat too loosely, described it as, “a one-stop shopping diagram that showed

how the Earth worked and how NASA’s missions to address this were going to satisfy the

science that needed to be done by the Bretherton Wiring Diagram. It was a very clever diagram

as a result.”113 One ESSC reviewer—atmospheric chemist Richard Stolarski—complained that

the ESS research program outlined by the ESSC reports and wiring diagrams was too inclusive:

“it seemed to have something for everyone at the expense of focussing on the critical problems

which block progress in understanding the system.”114 But this was the expressed and ultimate

intent of the ESSC, to incorporate the many different disciplines into one overarching rubric.

The reports and the Earth system wiring diagrams had to have something for everyone.

PROMOTING ESS

The Strategy

The ESSC took considerable measures to ensure that the broader scientific community was involved in its working groups and in the development of its reports. The ESSC also recognized

112 Chunglin Kwa, “Local Ecologies and Global Science: Discourses and Strategies of the International Geosphere- Biosphere Programme,” Social Studies of Science 35, no. 6 (Dec. 2005): 928. 113 Fisk, interviewed 8 Sep 2010. 114 Letter from Richard S. Stolarski to Francis Bretherton, 20 Apr 1987, Folder 48, Earth System Sciences Collection, NCAR, Boulder, CO.

256

the value of creating material that was aesthetically compelling, linking together all ESS

products into a recognizable brand, captured by the iconic wiring diagram of the Earth system.

This section focuses on the ESSC’s efforts to widely promote its ESS research program. The

effectiveness of this work with respect to the specific ESS research program was mixed at best

(chapter five), yet the sincerity and intensity of these efforts is readily apparent. The ESSC may

not have achieved its goal of instigating a new ESS research program, but its promotional efforts

contributed to the Earth system concept spreading to a variety of diverse constituencies, largely

among Earth scientists. ESS as a research program remained controversial due to its strong

emphasis on global scale research on time scales of decades to centuries using Earth observing

satellites. The program’s potential appetite for scarce research dollars meant that it was less

attractive to scientists working on longer timescales or those using in-situ observational techniques. The Earth system as a concept, however, captured the Earth science zeitgeist of the

1970s and 1980s, one that emphasized greater interdisciplinarity and regarded the Earth as an interconnected system of components. So long as there were no financial consequences attached to it, the Earth system concept was a useful interpretive tool for Earth scientists that flourished as a consequence of the ESSC’s promotional strategies (chapter five).

The ESSC undertook a number of activities in its effort to popularize the ESS research program. John McElroy and Ray Williamson note that the ESSC, “portrayed a picture of the

Earth sciences of remarkable breadth and great scientific attractiveness….NASA employed its considerable presentation skills in publishing widely read documents espousing the new view.”115 This sentence broadly and crudely describes the ESSC’s work, but it does not

sufficiently convey the ESSC’s efforts to “sell” its new research program. McElroy and William

115 McElroy and Williamson, “The Evolution of Earth Science Research from Space,” 449-50.

257

vastly underestimate the importance and extent of this work by limiting their focus to the ESSC’s

two published reports. They also do not recognize the important role played by non-NASA

personnel in these promotional activities. Though NASA personnel were certainly involved,

other actors like Payson Stevens, Laura Lee McCauley, and the non-NASA members of the

ESSC like Francis Bretherton and Berrien Moore all performed critical roles.

Efforts to spread the ESSC’s research program to a variety of constituencies was

spearheaded by Stevens and INI. It was McCauley, however, who initiated this conversation

amongst ESSC members. On December 27, 1985, McCauley sent a telemail to the ESSC

executive committee and Stevens regarding the “Promotion” of ESS:

Time to start thinking, and acting, on follow-up promotion of ESS through news releases, magazine stories...whatever (I don’t know much about the various avenues for exposure) to coincide with the release of the [Preview] brochure. I know Payson has had experience with this and maybe he could give us his thoughts. But we need to map out our strategy, etc. and have a central location for tracking it so the right stuff gets to the right people etc.116

This telemail began the first concerted effort to form a strategy to best present the ESS research

program to external communities. Bretherton credited McCauley for getting “this ball rolling”

and directed her to do “something about it[.]”117 This she did by reaching out to Stevens to draw

on his expertise in science communication and his earlier work on the Oceanography From

Space project. Stevens offered thoughts on “Formulating a Promotional Strategy for ESS” in

early January 1986. This strategy centered around the Overview report to be released in the

spring of 1986: “it is important to have an overall strategy for the release of information to

further ESS. What is needed (as they say in the business) is an orchestrated campaign which is

116 Telemail from L. McCauley to ESSC Executive Committee, 27 Dec 1985, Folder 55, Earth System Sciences Collection, NCAR, Boulder, CO. 117 Telemail from L. McCauley to ESSC Executive Committee, 7 Jan 1986, Folder 55, Earth System Sciences Collection, NCAR, Boulder, CO.

258

coordinated with key political events and reaches specific targeted audiences.” Stevens

emphasized that it was, “important that we move immediately to identify key dates throughout

1986 (and beyond) and coordinate specific media releases to promote ESS.”118

Stevens then provided a, “laundry list of potential products which can be considered

important elements for reaching a wide range of audiences” and that should all be thematically

and chromatically linked together as part of an ESS brand (see above). In addition to the

Preview and Overview reports, Stevens recommended developing: a folder containing “exciting

images”; postcards highlighting research and hardware that were, “beautiful enough to save and

pin up”; ESS posters (similar to those done for the Oceanography From Space project); a

convention display to “draw interest/recognition to [the ESS] program” at meetings like those held by the American Geophysical Union (AGU) and the American Association for the

Advancement of Science (AAAS); journalism pieces in science magazines (e.g. OMNI, National

Geographic), newspapers, airline magazines, and promotional releases sent out to journalists, “as potential storylines to whet their appetites”; a 10 to 15 minute video piece using “state-of-the-art

computer graphics”; and an audio-visual slide show.119 Not all of these materials came to

fruition. For instance, Stevens did not produce an ESS video piece, nor were postcards made.

Regardless, Stevens had a broad vision for selling ESS to experts and the public using a variety

of media.

Stevens did not simply suggest the development of materials for distribution, but also

made recommendations about how to communicate this material and ESS ideas more generally

to as many different communities as possible. These recommendations broadly broke down into

118 Telemail from P. Stevens to S. Tilford, et al., 3 Jan 1986, Folder 55, Earth System Sciences Collection, NCAR, Boulder, CO. 119 Ibid.

259

two main types. First, NASA and the ESSC could organize and initiate presentations and write-

ups on ESS. Stevens envisaged presentations to Congress, science organizations (e.g. UCAR),

international forums (e.g. ICSU) and at professional scientific meetings (e.g. AGU, AAAS).

Second, the ESSC must be ready to take advantage of more “random” opportunities that might

emerge from chance meetings with members of the White House, Congressional leaders, key

industry executives, NASA and NOAA administrators, and key international political and

scientific leaders. These happenstance meetings were, “not-to-be-missed chances to make a

major impact at a high level when people are uniquely receptive. Since these opportunities

cannot be predicted in advance we must be ready with a variety of materials in appropriate

media.” To facilitate this promotion, Stevens recommended the formation of a “Senior Advisory

Group for ESS” and listed potential celebrity candidates like Roger Revelle, Walter Sullivan

(New York Times science reporter), Paul Ehrlich, Garrett Hardin, Eugene Odum, Stephen Jay

Gould, and David Attenborough.120 Some of Stevens’ ideas, like the Senior Advisory Group,

might have been too ambitious for a small group of scientists and engineers formed under the

purview of NASA’s Advisory Council. However, many of Stevens’ suggestions were adopted,

enough to show that the ESSC clearly supported his sweeping plan to sell ESS.

Mailing Lists

Winter and Spring of 1986 represented a period of intense promotional and editorial activity for

the ESSC and its support staff. Not only was the ESSC deep in the last stages of editing its

interim report, the Overview, and incorporating comments from external reviewers, but it was

also formulating precisely how to unveil the ESS program to a range of different communities in

120 Telemail from P. Stevens to S. Tilford and L. McCauley, 9 Jan 1986, Folder 55, Earth System Sciences Collection, NCAR, Boulder, CO.

260

a way that would translate into concrete and broad-ranging support. What coalesced as the

official strategy was a combination of more informal conversations with coordinated dispersals

of the Preview brochure and Overview report, culminating in a formal press conference held at

NASA Headquarters in Washington, DC on June 26, 1986. This was just before the World Wide

Web and before email could move easily between different service providers. As a result, a

crucial component of the ESSC’s promotion strategy was the dissemination of material via

mailing lists. Historians such as Paul Edwards in The Closed World (1996) and David Kaiser in

How the Hippies Saved Physics (2011)121 note the importance of this mechanism for circulating

documents to build and maintain what computer pioneer Jay Forrester called a supportive

“outside constituency.”122

McCauley and other members of the ESSC project staff compiled comprehensive mailing

lists to distribute ESSC material, continually revising them as necessary. Back and forth

telemails between McCauley and other staff demonstrate the efforts taken to eliminate any

“gaps” in these lists for both the Overview and later Closer View distributions.123 Even with

these efforts, numerous individuals—both expert and lay—wrote to the ESSC requesting that

their names be added to its mailing list, or at minimum requesting copies of the ESSC’s reports.

These requests came from individuals from all over the world: Denmark, Japan, Canada,

Switzerland, the UK, Mexico, and China.124 Though it is not always clear how individuals first

121 Paul N. Edwards, The Closed World: Computers and the Politics of Discourse in Cold War America (Cambridge, MA: MIT Press, 1997); David Kaiser, How the Hippies Saved Physics: Science, Counterculture, and the Quantum Revival (New York: W.W. Norton, 2011). 122 Quoted in: Edwards, Closed World, 100. 123 Memorandum from Laura Lee McCauley to [ESSC] Members, et al., 3 Nov 1987, Folder 77, Earth System Sciences Collection, NCAR, Boulder, CO; Note from LL [McCauley] to “the OIES Gang,” 27 Aug [1987], Folder 78, Earth System Sciences Collection, NCAR, Boulder, CO. 124 Folder 74, Earth System Sciences Collection, NCAR, Boulder, CO; Letter from Payson Stevens to Panqin Chen, 20 Mar 1989, Folder 33, Earth System Sciences Collection, NCAR, Boulder, CO; Folder 58, Earth System Sciences Collection, NCAR, Boulder, CO; Folder 149, Earth System Sciences Collection, NCAR, Boulder, CO.

261

became aware of the ESSC’s work, it is possible that they may have come across an “ad” for the

Closer View report placed in various science publications. The ad briefly described the ESSC’s

work and provided a coupon that could be used to obtain a free copy of the report.125

Beyond the lists of individuals they amassed, ESSC staff also obtained mailing lists from

other organizations whose members might be receptive to the ESS program. One mailing list

alone, from NASA’s OSSA, contained around 10,000 names and addresses that the ESSC

merged into their own growing list.126 The ESSC also considered purchasing mailing lists from

the American Meteorological Society (around 7,000 names) and the Ecological Society of

America (around 10,000 names).127 This gives some sense of the magnitude of the ESSC’s

mailing outreach. Final prints for both the Overview and Closer View numbered in the tens of

thousands each.128 The amassed mailing lists contained thousands of names: scientists (e.g. H.T.

Odum, S. Fred Singer), members of science organizations (e.g. the AGU, AAAS, NAS), civil

servants (from NASA, NOAA, NSF, OSTP, OMB, DOD, and DOE), high school teachers and

university professors, industry representatives, politicians, congressional staffers (e.g. Radford

Byerly of the House of Representatives Committee on Science and Technology), journalists

(from Popular Mechanics, Science, CNN, United Press International, National Geographic), and

international contacts (from Austria, France, Saudi Arabia, Brazil, Japan, West Germany,

Australia, the United Kingdom, Italy, Canada, Indonesia, India, New Zealand, South Korea,

Belgium, Argentina, Switzerland, Mali, Netherlands, Nigeria, Thailand, Bangladesh, South

125 Suggested Text for Closer View Ad, n.d., Folder 77, Earth System Sciences Collection, NCAR, Boulder, CO. 126 Telemail from K. Wolfe to L. McCauley, 13 Feb 1986, Folder 32, Earth System Sciences Collection, NCAR, Boulder, CO. 127 Update on ESSC Activities and Plans, [24 Apr 1986], Folder 132, Earth System Sciences Collection, NCAR, Boulder, CO. 128 Ibid.

262

Africa, Pakistan, and Sweden).129 There were also numerous copies of the reports printed for

spontaneous circulation as ESSC members and others took advantage of what Stevens called the

“not-to-be-missed chances” to promote ESS.

Press Conference

If distributing reports to thousands of individuals was one of the most far-reaching activities

undertaken by the ESSC, its press conference was one of its most ostentatious. Held on June 26,

1986 at NASA Headquarters to publicly release the Overview report, this press conference was

the pinnacle of the ESSC’s promotional strategy. It combined report distribution with a variety

of other materials and techniques in an effort to create a lasting impression and widespread

interest. In this single event, most of Payson Stevens’ ideas for how to sell ESS were put to use.

Press conference planning began in the late winter and early spring of 1986. ESSC members

wanted the conference to have a “big splash” amongst scientists, politicians, and the public.130

This meant having all materials ready for distribution, ensuring the presentation room properly

showcased the ESS brand, making arrangements for the heads of NASA, NOAA, and NSF to

speak at the conference, and timing the conference so that members of Congress could attend. In

addition to relying again on the science communications savvy of Payson Stevens and INI, the

ESSC also worked with NASA Headquarters’ Public Affairs Officer Jim Kukowski. Kukowski

lent his NASA-earned public relations expertise to the ESSC, arranging everything from

129 Memorandum of Call, 5 Aug [1986], Folder 58, Earth System Sciences Collection, NCAR, Boulder, CO; Telemail from J. Dutton to L. McCauley, 21 Jan 1986, Folder 154, Earth System Sciences Collection, NCAR, Boulder, CO; Note, “To Receive Reports,” n.d., Folder 46, Earth System Sciences Collection, NCAR, Boulder, CO; Members of Press Who Have Expressed an Interest in ESSC, n.d., Folder 49, Earth System Sciences Collection, NCAR, Boulder, CO; Folder 59, Earth System Sciences Collection, NCAR, Boulder, CO; NASA International Agency List, ESSC and ESAD Annual Report Mailings, Folder 156, Earth System Sciences Collection, NCAR, Boulder, CO; Memorandum from Sarah Hawkins to Laura Lee McCauley, 14 Sep 1987, Folder 156, Earth System Sciences Collection, NCAR, Boulder, CO. 130 Telemail from N. Brewster to L. McCauley, 12 Jun 1986, Folder 50, Earth System Sciences Collection, NCAR, Boulder, CO.

263

mundane tasks like name tags, a glass display case, and an ESS button for Bretherton, to coordinating press releases and generating media interest by contacting news outlets like Time

Magazine, NBC, and USA Today.131 CNN actually showed the press conference live.

Certainly, the Challenger shuttle accident on January 28, 1986 contributed to a renewed public

interest in NASA, an interest that had been waning.132

The press conference was a “classy” affair. In keeping with ESS’s brand, McCauley

reportedly wore a “teal” suit to the event.133 For this event and others, Stevens designed an ESS

poster in “ESS blue” with the large ESS diamond logo dominating the majority of the page.

Superimposed onto the Earth at the center of the logo was the simplified version of the Earth

system wiring diagram. Surrounding the logo and diagram were images of a satellite, a forest, a volcano erupting, and satellite data superimposed on a map of Pacific Ocean wind directions.

Both the “goal” and “challenge” of ESS were stated at the bottom, along with a quotation from

T.S. Eliot: “We shall not cease from exploration/ And the end of all our exploring/ Will be to

arrive where we started/ And to know the place for the first time” (Figure 4.6).134 With broadly

recognizable images and a simple wiring diagram of the Earth system, the poster could promote

ESS to scientific experts, political leaders, and the general public. This and other evocative

images of the Earth and satellite data from the Overview were mounted on foam board adorning

131 Handwritten notes,[1986], Folder 32, Earth System Sciences Collection, NCAR, Boulder, CO; Handwritten notes, [1986], Folder 136, Earth System Sciences Collection, NCAR, Boulder, CO; Typed notes, [1986], Folder 50, Earth System Sciences Collection, NCAR, Boulder, CO. 132 Shuttle launches—including the ill-fated Challenger launch—had ceased to be televised live on national networks. The major exception here was CNN, which was the only national television channel to air the Challenger launch, and therefore the accident, live. See: Tricia Escobedo, “When a National Disaster Unfolded Live in 1986,” CNN, 31 Mar 2016, accessed 9 October 2018, https://www.cnn.com/2016/03/31/us/80s-cnn-challenger- coverage/index.html. 133 Telemail from N. Brewster to L. McCauley, 27 Jun 1986, Folder 137, Earth System Sciences Collection, NCAR, Boulder, CO. 134 “Payson R. Stevens: Earth System Science/ESS,” accessed 15 Jan 2020, https://www.paysonrstevens.com/science-communication/internetwork-incini/earth-system-science/.

264

the walls of the conference room.135 Press kits contained copies of the Preview brochure,

Overview report, a fact sheet, slide images, and a copy of Bretherton’s presentation (see below).

Documents were placed on display at the back of the room with “stacks below for the taking.”

The logos for ESS, NASA, NOAA, and NSF were placed “everywhere.”136

135 Telemail from P. Stevens to L. McCauley et al., 12 Jun 1986, Folder 50, Earth System Sciences Collection, NCAR, Boulder, CO; Art List, [26 Jun 1986], Folder 50, Earth System Sciences Collection, NCAR, Boulder, CO. 136 Telemail from L. McCauley to J. Baker et al., 23 Jun 1986, Folder 137, Earth System Sciences Collection, NCAR, Boulder, CO.

265

Figure 4.6. The Earth system science poster designed for the ESSC’s press conference, June 26, 1986. (Design by InterNetwork, Inc | Payson R. Stevens; reprinted with permission from Payson Stevens)

266

The press conference was invitation only. Those invited were from the most relevant

government agencies (e.g. NASA, NOAA, NSF, USGS, OSTP) and influential media outlets

(e.g. New York Times, Science, National Geographic), members of Congress, and prominent scientists. For instance, the ESSC made a special effort to invite famed atmospheric scientist

Bert Bolin, (who was in the US at that time), perhaps at the behest of Berrien Moore who had close connections to Bolin and viewed him as an important mentor.137 “Old friends of NASA”

like the author James Michener and atmospheric physicist Richard Goody (of Global Habitability

“fame”) also received invitations.138 The ESSC also used the press conference as an opportunity

to win over influential individuals who might potentially oppose the ESS research program.

Thomas Donahue, chair of the Space Science Board (SSB), was identified as, “a potential enemy

whom this [the ESSC press conference] might influence and convert to friendship; the SSB must

be made a friend of ESS.” Similarly, the National Academy of Sciences president Frank Press

was, “essential since he has been a reluctant father if not a foot-dragger.”139 These were

individuals whose support was viewed as crucial if ESS was to become an actual research

program. The ESSC hoped that the press conference could serve a broad consensus building

function, reaching beyond Earth scientists to target the larger scientific community, government

officials, and politicians. For the project to be a success, the ESSC needed to secure

endorsements from the upper echelons of the federal government and from the scientific elite.140

137 Telemail from J. Eddy to K. Holmen [for Bolin], 13 Jun 1986, Folder 136, Earth System Sciences Collection, NCAR, Boulder, CO; Moore interview, 4 Apr 2011. 138 Note on Press Conference Invitee Recommendations, [1986], Folder 50, Earth System Sciences Collection, NCAR, Boulder, CO. 139 [ESSC] Press Conference Invitee Recommendations, [1986], Folder 50, Earth System Sciences Collection, NCAR, Boulder, CO. 140 Though they didn’t directly control US government discretionary funding (that is one of the roles of Congress), federal agencies prepared initial budgets to present to the White House and OMB, which was the first step on the way to securing congressional funding.

267

Like many of the ESSC’s activities, conference preparations took longer than expected.

In part this was to ensure that the heads of NASA, NOAA, and NSF would all be available and

willing to endorse ESS publicly, which they all did. After the Challenger accident in January

1986, James Fletcher—who served as NASA’s Administrator from 1971 to 1977—was again

selected to head the agency during its difficult post-Challenger regrouping period. Fletcher

spoke at the ESS press conference. He drew on Apollo 8’s “Earthrise” image to compare the

planet to a “spacecraft” with a “life-support” system that required protection from the “deadly

vacuum” of space. Fletcher claimed that the ESS research program provided a strategy, “to

understand Earth as a system and improve the quality of life on our planet.” Alluding to the

changes being made “In the wake of the Challenger tragedy,” Fletcher noted that a major part of

NASA’s “regrouping” strategy was to use advances in satellite technology to study the planet in a more systematic way to better understand both natural and “man-made” global change. The

ESSC’s Overview report would, according to Fletcher, serve as a “blueprint” for NASA’s Earth science research, “through the 1990s and beyond.”141 Both NOAA’s Administrator (Anthony

Calio) and NSF’s Director (Erich Bloch) made similarly laudatory and supportive speeches,

clearly and publicly endorsing the ESS research program.142 Jack Eddy described this tri-agency

endorsement as “historic.”143 Perhaps not surprisingly, Bretherton gave the longest presentation

at the event, speaking for 20 minutes with a thick stack of vugraph slides prepared by Stevens

that outlined ESS with both engaging images and graphical details. It was followed by a 30-

141 Statement of Dr. James Fletcher, NASA Administrator, Earth System Science Press Conference, 26 Jun 1986, Folder 51, Earth System Sciences Collection, NCAR, Boulder, CO. 142 Remarks, Dr. Anthony Calio, Administrator, National Oceanic and Atmospheric Administration, Earth System Science Committee Report Briefing, 26 Jun 1986, Folder 51, Earth System Sciences Collection, NCAR, Boulder, CO; Statement by Mr. Erich Bloch, Director, National Science Foundation at the NASA Press Briefing, June 26, 1986, on the Occasion of the Release of the Report Overview of the Earth System Sciences Committee, 26 Jun 1986, Folder 51, Earth System Sciences Collection, NCAR, Boulder, CO. 143 Telemail from J. Eddy to K. Holmen [for Bolin], 13 Jun 1986, Folder 136, Earth System Sciences Collection, NCAR, Boulder, CO.

268

minute question and answer period, with some ESS “specialists” in the audience fielding specific

Earth science questions which Bretherton lacked the expertise to answer.144

In practical and immediate terms, what the ESSC envisioned for the press conference and

how it actually unfolded were a good match. Nancy Brewster of the NSF wrote to McCauley on

June 12 after the “pre-press briefing” held by NASA’s Kukowski: “It sounds like its [sic] really

going to be a big splash [with] all that they have planned.”145 McCauley agreed: “Yes I think the press conference will be a big hit[.]”146 Brewster’s expectations were not disappointed. She

wrote to McCauley the day after the conference, “to once more applaud your efforts. It couldn’t

have gone more smoothly or with a bigger bang. Everyone came together for the last hurrah and

I think everyone should be proud of doing ‘a good thing.’” Brewster reported that NSF Director

Bloch also, “felt it went well and was pleased with the whole thing.”147 McCauley herself

described the event as, “a MAJOR SUCCESS!!!”148 After months of preparations, everyone

agreed that they had been rewarded for their hard work. ESS received ample popular press

coverage and many new requests for its materials.149 But did the event actually convert those

more ambivalent about ESS and transform potential enemies into friends? Those strategic

results remain more difficult to judge. However, even if the ESSC’s primary and long-term strategic objectives were not fully met, the efforts the committee devoted to the press conference

144 Telemail from L. McCauley to J. Baker et al., 23 Jun 1986, Folder 137, Earth System Sciences Collection, NCAR, Boulder, CO; Typed notes, [1986], Folder 50, Earth System Sciences Collection, NCAR, Boulder, CO. 145 Telemail from N. Brewster to L. McCauley, 12 Jun 1986, Folder 50, Earth System Sciences Collection, NCAR, Boulder, CO. 146 Telemail from L. McCauley to N. Brewster, 13 Jun 1986, Folder 50, Earth System Sciences Collection, NCAR, Boulder, CO. 147 Telemail from N. Brewster to L. McCauley, 27 Jun 1986, Folder 137, Earth System Sciences Collection, NCAR, Boulder, CO. 148 Telemail from LL [McCauley] to Deb [Stirling], 27 Jun 1986, Folder 137, Earth System Sciences Collection, NCAR, Boulder, CO. 149 Telemail from N. Brewster to L. McCauley, 27 Jun 1986, Folder 137, Earth System Sciences Collection, NCAR, Boulder, CO.

269

demonstrate that the ESSC both recognized and acted upon the need to promote its fledgling ESS

research program.

Print Media

While the press conference represented the most conspicuous event produced by the ESSC, its efforts to promote ESS via other journalistic avenues arguably reached a wider audience. This was before PVRs, the World Wide Web, YouTube, and on-demand television. One-time events like a press conference might have received a certain amount of additional play time on, say, the evening news, but it was not something that could easily be viewed on call. Nor did a majority of people necessarily have access to the formal ESSC reports despite the extensive mailing lists.

Newspaper coverage and popular journal pieces that might be read “on demand,” therefore, served as another important avenue through which the ESSC promoted its research program.

Not surprisingly, NASA’s Kukowski ensured that there was press coverage of the press conference, and articles appeared sporadically after that, as journalists made links between the

ESSC’s work and environmental issues.150

Apart from these (largely) unsolicited pieces, ESSC members and affiliates also sought

out opportunities to publish popular pieces on ESS. Part of Payson Stevens’ ESS promotional

strategy from January 1986 involved coordinating the release of ESS stories in a variety of media

outlets (science magazines, newspapers, and airline magazines). Stevens had written for a

number of science magazines and had “good contacts.”151 Of these magazines, Stevens seems to

have gotten the farthest with OMNI, which was created by Penthouse publisher Bob Guccione

150 Folder 46, Earth System Sciences Collection, NCAR, Boulder, CO. 151 Telemail from P. Stevens to S. Tilford, et al., 3 Jan 1986, Folder 55, Earth System Sciences Collection, NCAR, Boulder, CO.

270

and Kathy Keeton in 1978. Patrick McCray describes OMNI as a “lush” publication that

capitalized on resurgent public interest in science and technology in the late 1970s and 1980s.

Future-oriented, OMNI presented a more optimistic view in which new technologies could

provide solutions to earthly social, environmental, and economic problems.152 Presumably an

Earth science research program that relied heavily on Earth observing satellites and numerical

modeling would be of interest to OMNI readers. The details of Stevens’ negotiations with OMNI

are not preserved, but on August 29, 1986 McCauley wondered “what’s happening” with the

OMNI story. Stevens had been busy with other jobs (“I do have other clients, believe it or

not?”).153 He eventually met with OMNI editor Patricia Aldrich on August 18. She expressed

interest, “but wants a strong hook for it.” Aldrich “stressed the need for tangible payoffs,

predictions and [the] role of satellites for their reader interest.” She suggested that a way to

encourage reader interest and participation was to include a “questionnaire” that would ask the

reader about the role of ESS and its place in NASA’s future. Stevens thought this could be “an

exciting element” but also potentially “dangerous if [the] results were unfavourable to ESS.”154

He added that McCauley and ESSC members should send him ideas and pictures to relay to

Aldrich. However, as of September 9, Stevens still had not “gotten any replies” to this request.155 Since there are no subsequent references in the archival documents nor is there an

ESS article in OMNI, this appears to have been the end of the attempt. While ESSC members

generally agreed that ESS promotion was essential, they did not always appear willing to

152 Patrick McCray, The Visioneers: How a Group of Elite Scientists Pursued Space Colonies, Nanotechnologies, and a Limitless Future (Princeton, NJ: Princeton University Press, 2013), 113-4. 153 Telemail from L.McCauley to P. Stevens, 29 Aug 1986, Folder 137, Earth System Sciences Collection, NCAR, Boulder, CO. 154 Telemail from P. Stevens to L. McCauley, et al., 30 Aug 1986, Folder 147, Earth System Sciences Collection, NCAR, Boulder, CO. 155 Telemail from P. Stevens to L. McCauley, 9 Sep 1986, Folder 6, Earth System Sciences Collection, NCAR, Boulder, CO.

271

contribute more of their time to that endeavor. This again points to the committee fatigue

described by Frederik Pohl’s “Speed Trap” (chapter three).

A somewhat more successful attempt to solicit a popular science article on ESS was

initiated by Berrien Moore. Moore had a cousin who worked at Time Magazine, and he managed

to generate “such enthusiasm” for ESS that Time’s science correspondent Dick Thompson agreed

to work with the ESSC to prepare a presentation to Thompson’s editors, “in hopes

(expectations?) that they will want to run a four-page spread on ESS!”156 On July 16, 1986

McCauley passed along instructions to the ESSC executive committee regarding this meeting to be held on July 22. Though it was scheduled for an hour and a half, it would be an informal affair with presentations occurring over lunch. Moore would be joined by Bretherton and Jim

Baker, and each would have two to four minutes to present, followed by a question and answer session. McCauley offered some warnings passed on by Thompson: “The editors will not have done any homework. This group is intimidated by academics and will be intimidated by anything technical--words like tropospheric chemistry, etc.” Thompson thought that linking ESS with “the greenhouse gas business” would be a “good lead in” since this topic has “been played up in the news lately[.]” Overall, the ESSC representatives, “Need to fire their imagination!!”157

There is no record for how this meeting went. The ESSC’s next documented interaction with

Time does not occur until September 1987, when McCauley wrote to Madeline Nash at Time

regarding her article on “global climate change.” After speaking with Nash, Bretherton had

requested that McCauley forward Nash some ESSC documents (the Preview and Overview,

along with a copy of the Closer View that had “just gone to press”). While Time did not run the

156 Telemail from LL [McCauley], [1986], Folder 137, Earth System Sciences Collection, NCAR, Boulder, CO. 157 Telemail from L. McCauley to ESSC Executive Committee, 16 Jul 1986, Folder 137, Earth System Sciences Collection, NCAR, Boulder, CO.

272

“four-page spread on ESS” as originally hoped, nor is it clear that it ran a piece on global climate

change that mentioned ESS in 1989, it did change its “Person of the Year” to “Planet of the

Year” in 1989, with “Endangered Earth” taking home the honour. Dixon Butler called Moore a

“political player” and credited him with this achievement: “Berrien made that happen. He has

contacts I just couldn’t believe in those days.”158

Presentations, Education Outreach, and International Liaisons

Another promotional strategy was to present ESS material at any and every opportunity, be it at

an academic conference or scientific meeting or special interest group or to congressional

staffers. Though a number of ESSC members would make these presentations, Francis

Bretherton gave the majority of them. Bretherton presented to a number of different groups:

NASA officials, members of NASA’s Advisory Council, congressional staff, Congress, the

IGBP, and the ICSU.159 As early as August 1985, before the ESSC had published a single

document, Bretherton even gave a presentation to the National Commission on Space.160

President Ronald Reagan established this group in 1984 to discover and evaluate future

directions for the US space program. It included prominent members like former NASA

Administrator Thomas Paine, Neil Armstrong, space colony enthusiast Gerard O’Neill, and

sound-barrier-breaking test pilot Chuck Yeager.161 McCauley, Blanchard, and Stevens

158 Dixon M. Butler, interviewed by Rebecca Wright, Washington, DC, 25 Jun 2009, NASA Johnson Space Center “Earth System Science at 20 Oral History Project,” accessed 10 Oct 2019, https://historycollection.jsc.nasa.gov/JSCHistoryPortal/history/oral_histories/NASA_HQ/ESS/ButlerDM/ButlerDM _6-25-09.pdf. 159 Memorandum and Attachments from Laura Lee McCauley to ESSC Members, et al., 15 May 1986, Folder 71, Earth System Sciences Collection, NCAR, Boulder, CO. 160 Presentation by Francis Bretherton on Earth System Sciences Program Strategy to the National Commission on Space, 22 Aug 1985, Folder 100, Earth System Sciences Collection, NCAR, Boulder, CO. 161 “Executive Order 12545 -- National Commission on Space,” Ronald Reagan Presidential Library, accessed 27 Oct 2019, https://www.reaganlibrary.gov/research/speeches/11486b; “Appointment of 14 Members of the National

273

developed a standardized, “set of material for visual presentations of the ESSC conclusions-- including background, introduction, science rationale, program recommendations, agency roles/budget, and other issues, plus some of the images[.]” All of the slides had uniform borders branded by the recognizable “ESS blue” and the ESS logo. These were distributed to every

ESSC member for use.162 Using these materials, the ESSC regularly set up displays at AGU’s fall and spring meetings.163 However, it occasionally participated in greater depth. For instance,

at the AGU’s spring meeting in 1985, the ESSC organized an entire panel on ESS in which

members presented the project from the vantage point of their different areas of expertise.

Chaired by atmospheric scientist and ESSC member Ronald Prinn, it was comprised of six

presentations by other ESSC members on components of the Earth system.164

NASA personnel also gave ESS presentations, as attested by Diane Wickland, a

terrestrial ecologist and NASA scientist since 1983. Wickland reported that NASA public

outreach was quite intense at the time:

We briefed anybody and everybody about Earth System Science, EOS, a particular aspect of it, whatever. Back in those days we didn’t have [Microsoft] PowerPoint files and computers, we had these transparencies called viewgraphs. We had probably 50 different presentations customized to different aspects of the program, as well as a couple different broad overviews….The outreach was very active. We touched a lot of audiences. Any scientific meeting you could get a paper on the agenda, somebody did. We just spread the word, created interest.165

Commission on Space, and Designation of the Chairman and Vice Chairman,” Ronald Reagan Presidential Library, accessed 27 Oct 2019, https://www.reaganlibrary.gov/research/speeches/32985a. 162 Letter from Laura Lee McCauley to Earth System Sciences Committee Members, 31 May 1986, Folder 29, Earth System Sciences Collection, NCAR, Boulder, CO. 163 Memorandum and Attachments from Laura Lee McCauley to ESSC Members, et al., 15 May 1986, Folder 71, Earth System Sciences Collection, NCAR, Boulder, CO. 164 AGU Presentations, 29 May 1985, Folder 136, Earth System Sciences Collection, NCAR, Boulder, CO. 165 Diane E. Wickland, interviewed by Rebecca Wright, Washington, DC, 26 Mar 2010, NASA Johnson Space Center “Earth System Science at 20 Oral History Project,” accessed 27 Oct 2019, https://historycollection.jsc.nasa.gov/JSCHistoryPortal/history/oral_histories/NASA_HQ/ESS/WicklandDE/Wickla ndDE_3-26-10.pdf.

274

These kinds of presentations were an effective means of spreading awareness of ESS, even if

they did not always have the intended effect of generating support for the ESS research program.

In addition to simply raising awareness of the ESS program in scientific and political

communities, the ESSC also sought to capture the imaginations of students not-yet-established in research fields or interest communities. A critical component of the ESSC’s promotional strategy involved educational outreach and the creation of ESS opportunities for post-secondary and post-doctoral students. Education would be important to familiarize students (and perhaps more seasoned practitioners) with the interpretation and use of remote sensing data for Earth science research, something that Bretherton and others noted was still in its infancy in the mid-

1980s. According to Bretherton, “even with the enormous strides that have already been made, the use of remote sensing in the Earth Sciences is still in its infancy. In almost all fields it is still a minority technique compared to in-situ observations, and the training of experienced practitioners knowledgeable across the full range of skills required is a major long-term task.”166

Though the ESSC claimed, in its Closer View report, that much of the “skilled and innovative

personnel” required for an ESS research program will come from “existing communities,” it

argued that, “in the longer term the initiative must develop its own organized body of knowledge

and recognized educational pathways.” Most importantly, these ESS educational pathways

would develop in universities, which should, “provide an emphasis that makes the endeavor

visible and attractive. Faculty members engaged in Earth system research must take the

initiative to lead their graduate students in the same direction.” The report also recommended

166 Francis P. Bretherton, “Earth System Science and Remote Sensing,” Proceedings of the Institute of Electrical and Electronics Engineers 73, No. 6 (June 1985): 1122.

275

creating ESS fellowships to increase graduate student “motivation.”167 This would generate a

continual crop of ESS-interested scientists to expand the research program.

Lastly, the ESSC placed high importance on linking ESS to other international projects

and organizations. Indeed, “international cooperation” was one of the ESSC’s main

recommendations. According to the ESSC, studying the Earth as a system was an “inherently

international” enterprise since it focused on global changes and would require not just space-

based but also in-situ measurements taken from (eventually) all countries.168 However, it was

also a way to justify ESS by placing it within larger international Earth science trends and

preventing the kind of international backlash experienced by an ESS precursor, Global

Habitability. International programs complemented by ESS or to which ESS was relevant

included those conducted or planned by the WCRP (e.g. the Tropical Ocean Global Atmosphere

Program, WOCE, the International Satellite Cloud Climatology Project) and the International

Council of Scientific Unions (e.g. the International Lithosphere Programme, the Global Ocean

Flux Study). The most important one was certainly the ICSU’s International Geosphere-

Biosphere Program (IGBP). Officially launched in 1987, the IGBP coordinated research on

global and regional changes affecting the Earth’s physical, chemical, and biological systems.

While much of the primary impetus for the IGBP came from within the US—and indeed from

key actors involved in the ESSC—it was established by an international institution and had active participation from many so-called developing countries.169 As mentioned above,

167 ESSC, Closer View, 172. 168 Ibid, prologue. 169 Wickland interview, 26 Mar 2010.

276

Bretherton and other ESSC members continually presented ESS concepts and recommendations

at IGBP meetings and made sure to align their activities as closely as possible.170

CONCLUSION

The ESSC endeavoured to develop and present its ESS research program to a large number of

constituencies using a variety of materials and techniques. Members recognized that they were

proposing something so sizable, new, and expensive that it would require widespread support

and a massive consensus-building effort. The group’s efforts can be broken into three broad

components. First, the ESSC conducted extensive internal and external review processes for

their reports that incorporated the views of hundreds of different scientists and engineers.

Second, the ESSC placed high value on making their products visually compelling. The ESSC

hired Payson Steven’s graphic design company INI to promote ESS by creating an identifiable

“ESS” brand. The ESSC also created wiring diagrams that served, in part, as visual heuristic

devices for conscripting individuals into the ESS research program. Lastly, the ESSC embarked

on an expansive promotional strategy for ESS. This included a massive report mailing

campaign, a press conference, journalistic pieces, presentations, educational support, and

establishing links with related international programs already underway.

All this suggests that the ESSC was well aware of the need to build consensus among

scientists, politicians, government officials, and the general public about the importance of its

ESS research program. It sets the ESSC apart from similar contemporary committees, none of

which attempted such an expansive promotional campaign. That the specific ESS research

program never achieved a widespread consensus was not from any lack of effort on the ESSC’s

170 Telemail from L. McCauley to S. Tilford, et al., 5 Feb 1986, Folder 38, Earth System Sciences Collection, NCAR, Boulder, CO; Handwritten notes, n.d., Folder 100, Earth System Sciences Collection, NCAR, Boulder, CO.

277

part. Rather, the inherent tensions in the specific ESS program—that prioritized certain kinds of

Earth science research over others, and therefore marginalized certain groups—created obstacles

that were difficult to overcome with so much potential funding at stake. No amount of “selling”

the ESS research program could convince a geophysicist that the Earth system should be defined

in the way chosen by the ESSC, which limited the Earth system to that bounded by the surface of

the planet and the mesopause. No amount of selling could convince, say, a seismographer that

the Earth system should be defined in a way that marginalized research on the Earth’s core. For

a geophysicist to agree with the ESSC would be to agree that research that focused on longer

term Earth processes should be deprioritized in favour of research conducted by those working in

fields that focused on shorter term processes (e.g. those working in atmospheric science,

oceanography, and ecology). By defining the boundaries of the Earth system in a way that

placed much geophysical research at the periphery of an ESS research program, the ESSC

generated apprehension in geophysics communities. Geophysicists largely rejected the specific

ESS research plan. However, they readily adopted the vaguer notion of the Earth system. They

simply chose to define the Earth system in a more inclusive way, in a way that incorporated

geophysical research interests. Though the ESSC failed in establishing an ESS research

program, the Earth system concept that it developed and promoted as part of the program gained

extensive traction. The Earth system concept was vague enough to capture the general

sentiments of many Earth scientists—growing since the 1970s—that the Earth was a single, interconnected system, while allowing specialists to define the Earth system in different ways

(chapter five). The Earth system concept did this without being linked to what some viewed as a problematic research program.

Chapter 5 The Strength of Vagueness: The AGU and the Spread of the Earth System

INTRODUCTION

Popularizing the phrase “Earth system” was neither the ESSC’s “goal” nor its “challenge.” The

primary aim of the ESSC was to develop and promote a large-scale Earth science research program called Earth system science (ESS) that would link together practitioners in the various

Earth sciences using satellite observations and in-situ studies to collect data and better model the interconnected processes that constituted the Earth system (chapters three and four). In other words, the aim of the ESSC was to develop and sell ESS, a research program that focused on studying the Earth as a system. This new research program, with its specific implementation strategy, was not without its detractors. Many Earth science communities, notably the solid

Earth sciences, felt marginalized by ESS’s strong emphasis on decadal timescales over longer term changes, as well as the seeming prioritization of space observations over in-situ studies.

This marginalization would have real-world, funding implications if ESS were implemented, so

ESS was not merely ignored but often actively critiqued. Solid Earth scientists frequently resisted ESS as a specific research program, despite all of the promotional efforts undertaken by the ESSC to engender widespread support.

In the process of the ESS promotional efforts, ESSC members achieved an unintended result. While advocating for ESS, the ESSC concomitantly, though unwittingly, spread the use of a singular phrase—the Earth system—that could fill the growing semantic void among Earth scientists. This phrase referred to the planet as a system of interconnecting subsystems that required interdisciplinary Earth science research in order to properly (or at least better) understand it. The Earth system phrase proved useful to various Earth scientists not because it

278 279

meant something narrow and specific but rather because it was vague. In its details, the Earth system meant different things to different specialists. But in its broadest sense, the Earth system referred to something that all Earth scientists could agree about. Imprecision, therefore, allowed specialists to cross disciplinary boundaries to communicate and work with other specialists who also recognized the need to move beyond compartmentalization in the Earth sciences to develop a better scientific understanding of the entire planet. They could make interdisciplinary

connections without needing to agree on the specifics of what, exactly, the Earth system was.

Earth scientists agreed that there was an Earth system, even if they did not agree on the exact

details about what made it up. The ESSC had begun using the Earth system concept to facilitate

interdisciplinary cooperation among the Earth sciences. Recall that the Earth system wiring

diagram was used to point out where different disciplines “fit” into ESS. The Earth system

provided a common language and understanding of the planet without necessarily supporting the

ESSC’s specific research program that some scientists found problematic. Once released into the

“wild” with the promotion of ESS, the Earth system can be thought of as a “boundary object”

that Earth scientists used to communicate across disciplinary boundaries.

Drawing on scholarship in the history and sociology of science, chapter five first

examines the usefulness of vagueness and the concept of a boundary object. Initially developed

by Susan Leigh Star, a boundary object is something that specialists use in narrowly prescribed

ways, but that has enough interpretive flexibility that it can transcend disciplinary boundaries and

promote interdisciplinary collaboration despite a lack of consensus about specifics.1 In line with

Star’s position that boundary objects can be conceptual “objects,” this chapter argues that the

1 Susan Leigh Star, “The Structure of Ill-Structured Solutions: Boundary Objects and Heterogeneous Distributed Problem Solving,” in Readings in Distributed Artificial Intelligence, eds. Alan H. Bond and Les Gasser (San Mateo, CA: Morgan Kaufmann Publishers, 1988), 37–54.

280

Earth system concept, manifested in the simplified Earth system wiring diagram, may be

understood as a boundary object.

The second part of the chapter uses the case study of the American Geophysical Union

(AGU) to trace the movement of the Earth system concept beyond the ESSC. Self-proclaimed as

the largest US scientific society devoted to studying the Earth and its environment in space, by

1989 the AGU’s membership numbered over 20,000 and its research interests traversed Earth

and space science disciplines.2 With its massive size and interdisciplinary focus, the AGU

provides a critical case study of how the Earth system concept spread quickly beyond the ESSC

in the mid- to late-1980s, even while its ESS research program never fully materialized. It shows

the efficaciousness of the ESSC’s promotional efforts even if it did not fully achieve its intended

objectives. The case study further shows how knowledge and ideas flow between different

practitioners and institutions and scientific communities. Recent scholarship—notably John

Krige’s edited volume How Knowledge Moves (2019)—emphasizes the constraints that impede this flow and the difficulties involved in actually constructing a successful knowledge network.3

While the movement of knowledge is undeniably difficult and takes a vast mobilization of

people, resources, political will, technical capabilities, and a variety of other factors, it does

happen. There is what Bruno Latour calls, “the slight surprise of action.”4 The movement of the

Earth system concept beyond the ESSC to the broader Earth science community is this kind of

“success” story. Using AGU archival documents, this case study illustrates how the Earth

2 Letter from Don L. Anderson to George Bush, 17 Jan 1989, Folder 10, Box 30, AGU Records, American Institute of Physics (AIP), College Park, MD; Ronald E. Doel, “American Geophysical Union,” in History of Science in the United States, ed. Marc Rothenberg (New York: Garland Publishing Inc., 2001), 30. 3 John Krige, ed., How Knowledge Moves: Writing the Transnational History of Science and Technology (Chicago: University of Chicago Press, 2019). 4 Bruno Latour, Pandora’s Hope: Essays on the Reality of Science Studies (Cambridge, MA: Harvard University Press, 1999), ch. 9.

281

system concept was adopted by groups external to the ESSC, even while the ESS research

program did not gain such traction.5

THE UTILITY OF SEMANTIC VAGUENESS

In August 2006, 420 members of the International Astronomical Union (IAU) voted on

Resolution 5A, the definition of a “planet.” After the vote, a celestial body would be called a planet if and only if it was in orbit around the sun, was round in shape, and had cleared the area around its orbit of debris.6 With the passing of this resolution, the IAU formalized what

previously had been only an informal designation. Previously, dynamicists and structuralist

astronomers, amateurs, and the general public all used “planet” in different ways. Despite the

variety, Lisa Messeri argues that, before the IAU’s vote, the meaning of “planet” was similar

enough amongst the distinct groups that they could communicate without confusion, but

contained enough interpretive flexibility for there to be different meanings among different local

groups. This changed in August 2006 when the IAU voted to specify the definition of a planet.

Up until this point, the meaning of “planet” was vague. After this point, there were stipulated

necessary and sufficient conditions for membership in the category “planet.” All vagueness had

been removed. Messeri argues that semantic and technical clarity came at the expense of broader

utility, removing a key communication tool for the different social groups. Vagueness, as the

Pluto case study shows, can actually be quite useful for collaboration and communication. It can

5 The AGU archival material is housed at the AIP in College Park, Maryland. 6 International Astronomical Union, “IAU 2006 General Assembly: Result of the IAU Resolution Votes,” accessed 25 May 2019, https://www.iau.org/static/archives/releases/doc/iau0603.doc.

282 facilitate the spread of concepts since different users need not fully agree on the specific, local meaning of a particular word, phrase, or concept.7

At least since the logical positivists’ early-twentieth century call for a “precisification” of scientific terms, vagueness has often been thought of as a vice in the sciences.8 However, Susan

Leigh Star’s “boundary object” helps explain the way vagueness is sometimes useful in the sciences by facilitating interdisciplinary communication in the face of local differences in meanings.9 Star, along with collaborators James Griesemer and Geoffrey Bowker, developed the concept in response to Star’s sociological fieldwork among scientists.10 Despite lacking certain techniques that might aid in cooperative efforts—good models of other scientists’ work, shared units of analysis, methods for collecting and aggregating data, common goals or timescales—

Star discovered that scientists could, and often did, coordinate and communicate in cooperative situations without agreeing at the local, specialist level.11 They did so, in part, by making use of what Star calls “boundary objects.” Boundary objects are, according to Star and Griesemer,

“both plastic enough to adapt to local needs and the constraints of the several parties employing

7 Lisa Messeri, “The Problem with Pluto: Conflicting Cosmologies and the Classification of Planets,” Social Studies of Science 40, no. 2 (Apr. 2010): 187-214. 8 See: Rudolph Carnap, “The Elimination of Metaphysics Through Logical Analysis of Language,” in Logical Positivism, ed. A.J. Ayer (New York: The Free Press, 1959), 60-81. 9 Messeri relies predominantly on Peter Galison’s “trading zone” concept to argue for the utility of vagueness rather than “boundary object,” though the latter is more useful for this dissertation. Galison uses the “trading zone” metaphor to describe the diversity of scientific and engineering cultures in US particle physics communities throughout the twentieth century and the tensions that persisted between the autonomous and interconnected activities of these groups. Galison explains that coordination could occur “without homogenization” via “contact languages” that facilitated local coordination in support of a common objective without requiring globally agreed upon meanings for classification schemes, significations, or standards. While similar to a “boundary object,” Galison’s trading zone functions in reverse, whereby it is local activities that are coordinated while globally agreed upon meanings remain elusive. Since the Earth system concept functions more like a boundary object than a Galisonian trading zone or contact language, the dissertation relies on the former rather than the latter conceptual apparatus. See: Peter Galison, Image and Logic: A Material Culture of Microphysics (Chicago: University of Chicago Press, 1997), 783–4, 803. 10 Star, “The Structure of Ill-Structured Solutions,”; Susan Leigh Star and James R. Griesemer, “Institutional Ecology, ‘Translations’ and Boundary Objects: Amateurs and Professionals in Berkeley's Museum of Vertebrate Zoology,” Social Studies of Science 19, no. 3 (Aug. 1989): 387–420; Geoffrey C. Bowker and Susan Leigh Star, Sorting Things Out: Classification and Its Consequences (Cambridge, MA: MIT Press, 1999). 11 Star, “The Structure of Ill-Structured Solutions,” 46.

283

them, yet robust enough to maintain a common identity across sites. They are weakly structured

in common use, and become strongly structured in individual-site use.”12 Scientists “tack”

between the more structured (specialist) and less structured (interdisciplinary) meanings of

boundary objects, depending on the context.13 This variety of meanings—this vagueness—

allows scientists in different disciplines to cooperate in interdisciplinary settings without

necessitating total consensus at the disciplinary level.

Like the boundary objects themselves, Star’s concept has proved useful for historians,

sociologists, anthropologists, and other “science studiers,” including those studying the Earth and

environmental sciences. Paul Edwards applies the concept to both climate models and global

datasets to help understand how scientific and political discourses on the climate shifted from

more local to global concerns by the 1980s, transitioning from a “climatology” focused on

geographically-specific datasets to a model-based, temporally-dynamic, and globally-oriented

“climate science.” Computer models—particularly GCMs—and global datasets served as fundamental organizing principles that traversed the boundaries of disciplines like meteorology, oceanography, and ecology to create new scientific collaborations focused on constructing climate science knowledge.14 Chunglin Kwa analyzes the “interdisciplinarity” of the IGBP as a

boundary object, though Kwa argues that this boundary object was imposed on a group of

practitioners rather than emerging organically. The interdisciplinarity of the IGBP centered

around global datasets produced by Earth observing satellites. Kwa argues that in the 1980s

ecologists were encouraged to alter their studies from local plant and animal communities to

12 Star and Griesemer, “Institutional Ecology,” 393. 13 Susan Leigh Star, “This is Not a Boundary Object: Reflections on the Origin of a Concept,” Science, Technology, & Human Values 35, no. 5 (2010): 604–5. 14 Paul N. Edwards, “Representing the Global Atmosphere: Computer Models, Data, and Knowledge about Climate Change,” in Changing the Atmosphere: Expert Knowledge and Environmental Governance, eds. Clark A. Miller, Paul N. Edwards (Cambridge, MA: MIT Press): 31–65.

284

target the “global earth system” by NASA and other institutions in the name of

interdisciplinarity. Tensions emerged as different ecologists established their own particular

interpretations of the “global” that would incorporate previous scales of ecological research

while still participating in nascent interdisciplinary Earth and environmental science research

linked to the IGBP.15 Steve Easterbrook alters this analysis slightly, arguing that climate

models—manifested as lines of explicit and unambiguous computer code—are not boundary

objects since they allow no interpretive flexibility. However, Easterbrook maintains that specific

concepts in climate science—like “climate sensitivity”16 and even the concept of “model” itself—do serve as boundary objects. Both concepts possess a plasticity that allows for different local interpretations while still maintaining enough shared meaning to facilitate communication between practitioners in different disciplines and institutions.17

The Earth System as a Boundary Object

Despite disagreements among disciplines about its precise definition, the Earth system’s ubiquity today suggests that the concept has broad utility and applicability in both interdisciplinary and specialist contexts. It can, therefore, be thought of as a boundary object. In its broadest sense,

Earth scientists readily use the “Earth system” to refer to an interconnected planet that should be studied as a system via interdisciplinary and global research. However, it was also malleable enough to allow interpretive flexibility at the local level. Different Earth science disciplines could, and still do, define the boundaries of the Earth system in different ways (chapter three).

15 Chunglin Kwa, “Local Ecologies and Global Science: Discourses and Strategies of the International Geosphere- Biosphere Programme,” Social Studies of Science 35, no. 6 (Dec. 2005): 923–50. 16 Climate sensitivity refers to the amount of warming that will occur if the level of atmospheric carbon dioxide was instantaneously doubled from pre-industrial levels (i.e. from 280 to 560 ppm). See: Andrew E. Dessler, Introduction to Modern Climate Change (New York: Cambridge University Press, 2012), 99. 17 Steve Easterbrook, “Climate Models as Boundary Objects,” lecture, Issues in the Theoretical Foundations of Climate Science: Scientific and Philosophical Perspectives, Toronto, CA, 15 Nov 2018.

285

This multiplicity of meanings—this vagueness—allows Earth scientists to “tack” between a more strongly structured meaning in disciplinary settings and a less structured meaning in interdisciplinary settings. As it was developed and promoted by the Earth System Sciences

Committee, the “Earth system” maintained a semantic flexibility that precipitated its rapid spread and adoption, and helps account for its ubiquity today. The particular components of what constitutes the Earth system might be disputed, but that there is an Earth system is not questioned by Earth scientists.

The most lasting product of the ESSC is likely its conceptual model of the Earth system depicting processes occurring on timescales of decades to centuries, the wiring diagram that came to be popularly known as the Bretherton diagram. In fact, the ESSC produced two of these

Earth system diagrams, a more simplified and a more complex version (Figure 4.5, Figure 3.1).18

ESSC members viewed the more complex of these as the scientific centerpiece of their work.

Though not complete, the committee believed that it was a more detailed and accurate

representation of the Earth system that could serve as a base for future elaboration by ESS

research. ESSC members also believed that the diagram could serve as a consensus building tool

for their ESS research program (chapter four). Presumably, more detail would help Earth

scientists to more concretely situate their research into the overarching ESS framework.19

Despite the ESSC’s emphasis on the more complex diagram, it was the simplified version of the

Earth system wiring diagram that gained popularity and longevity. When scientists (and others)

today refer to the Bretherton diagram, they generally mean the simplified wiring diagram.

18 Earth System Sciences Committee (ESSC), Earth System Science: A Program For Global Change: A Closer View (Washington, DC: NASA, 1988). 19 Edward S. Goldstein, “NASA’s Earth Science Program: The Bureaucratic Struggles of the Space Agency’s Mission to Planet Earth” (PhD dissertation, George Washington University, 2007), 134.

286

That the simplified diagram became the canonical representation of the ESSC’s work

suggests that it was more useful for Earth scientists than the complex one. The “Earth system”

represented in the simplified wiring diagram is an “ideal type” of boundary object. Star and

Griesemer describe an ideal type as: “an object such as a diagram, atlas or other description

which in fact does not accurately describe the details of any one locality or thing. It is abstracted

from all domains, and may be fairly vague. However, it is adaptable to a local site precisely

because it is fairly vague; it serves as a means of communicating and cooperating symbolically - a ‘good enough’ road map for all parties.”20 They use the example of “species,” a concept that

describes no individual specimen but incorporates concrete and theoretical data into something

that helped communication among ecologists, amateur collectors, and hunters at Berkeley’s

Museum of Vertebrate Zoology.21

The utility of the simplified wiring diagram did not result from its accurate depiction of

an interconnected planet but from its depiction that the planet was interconnected. Though it

arguably contained more scientific realism, the complex wiring diagram was too intimately

linked with the ESS research program. It depicted a specific program where certain specialists

would be marginalized while others prioritized based on their areas of research. Ironically, it

was the attempt at realism in the complex wiring diagram that reduced its usefulness as a

boundary object. The complex diagram was too specific about the subsystems to be modelled to

be useful for different Earth science disciplines that wanted to engage with the Earth system in

different ways. Because it more closely depicted the specific ESS research program, the

complex diagram risked alienating certain Earth scientists because it depicted only shorter term

20 Star and Griesemer, “Institutional Ecology,” 410 [emphasis added]. 21 Ibid, 410.

287

processes as being internal to the Earth system, emphasizing this research at the expense of

longer term processes.

The simplified diagram, on the other hand, was too general to be threatening, alienating,

or confusing. It clearly showed, in broad strokes and at a single glance, that the Earth could be

understood as an interconnected system comprised of components studied by the various Earth

and environmental science disciplines. The simplified Earth system wiring diagram was not an

exercise in “scientific realism” (whatever that might look like), but instead was a very general

and abstract representation that facilitated interdisciplinary communication while allowing local

expert communities to define the boundaries of the Earth system much more narrowly. Though

it still depicted solid Earth and solar-terrestrial processes as being only “external” inputs to the system, this mattered little since the visual did not imply any specific research program or funding priorities. It was just a broad conception of the planet shared among Earth scientists.

While the Bretherton diagram may have largely failed as a consensus building tool for the

ESSC’s specific ESS research program, it succeeded in another way. It was a crisp visualization of a general idea nascent in the Earth and environmental sciences, that the Earth is a system made up of interconnected subsystems. This was the boundary object that facilitated communication and cooperation amongst Earth and environmental scientists with diverse disciplinary concerns.

THE AGU AND THE EARTH SYSTEM

Chapter three examined some of the obstacles the ESSC encountered when constructing its ESS research program, in particular the disagreements with solid Earth scientists. Though the ESSC eventually incorporated an Earth system wiring diagram that detailed processes occurring on timescales of thousands to millions of years (Figure 3.2), ultimately the reservations of solid

288

Earth scientists did not shift the overall central emphasis of ESS beyond the shorter timescales.22

As a consequence, solid Earth scientists responded less than enthusiastically to the ESSC’s specific ESS research program. Despite this, the Earth system concept proved popular among geophysicists in particular, who were quick to adopt the phrase to describe the Earth. An examination of the activities of the American Geophysical Union (AGU) in the 1980s helps show how the ESSC’s ideas—namely the Earth system concept—moved quickly and easily beyond the confines of the ESSC, despite the controversy and concerns raised about the ESS research program. The AGU’s large size and interdisciplinary character makes it a useful case study to help understand the spread of the Earth system concept. AGU members desired a research program that studied the Earth as a system, but they viewed ESS as being too narrowly focused and they did not want to marginalize research on longer term processes. They sought a research program that defined the Earth system differently than the ESSC. For the AGU, the

Earth system was not delineated by the lithosphere and the mesopause, but rather extended all the way down to the center of the planet and up to the outer reaches of the magnetosphere. The

AGU formed committees to develop a more comprehensive framework for studying the whole

Earth, what it called a “Complete Mission to Planet Earth.”23

Not only was the AGU one of the largest scientific organizations in the US, but it has a long history of interdisciplinarity. This traces back to the roots of geophysics itself. Geophysics emerged as a discipline distinct from geology in the late nineteenth century. Rather than sustaining geologists’ narrow concern for local descriptions of the Earth’s surface layer or to studying various parts of the planet in isolation, Peter Bowler argues that geophysicists sought to

22 ESSC, Closer View, 26. 23 A Complete “Mission to Planet Earth”: A Policy Statement of the American Geophysical Union, December 1989, 4 Nov 1989, Folder 10, Box 20, AGU Records, AIP, College Park, MD.

289

understand, “the physical properties of the earth as a whole.” Geophysics was, from the start,

“designed to synthesize knowledge of all the physical processes operating within the earth’s

crust, its oceans and its atmosphere.”24 The National Academy of Sciences (NAS) founded the

AGU in 1919 as an executive committee of the National Research Council (NRC) to provide a

professional society for studying the physics of the Earth. It became an independent

organization in 1972. As defined by the AGU’s statutes, the union’s broad purposes were to

promote the scientific study of the Earth and its environs, promote cooperation with other

geophysical scientific organizations, convey geophysical research results to other scientists and

the public (including politicians) via sponsorship of meetings and various publications, attract

students to geophysics, and encourage “new relationships” among scientists working in

geophysical-related disciplines.25 Ronald Doel notes that the AGU gave priority to “pure”

scientific research over “exploration” geophysics that developed techniques for discovering

petroleum deposits. By the 1950s, the AGU promoted research aligned with an expansive

interpretation of geophysical concerns, including meteorology, oceanography, geodesy,

seismology, terrestrial magnetism and electricity, volcanology, tectonophysics, and hydrology.

Space science joined the entourage of disciplines in the 1960s, with interest in ecology and other environmental sciences being a more recent development (see below).26

The AGU promoted research not only on the Earth but also the Earth’s environment in

space and, by the mid-twentieth century, it included research on other planets in the solar

24 Peter J. Bowler, The Norton History of the Environmental Sciences (New York: W.W. Norton and Company, 1992), 393. Bowler cites the example of Alfred Wegener, an early proponent of continental drift, who worked in both meteorology and geophysics. According to Bowler, Wegener, “was an exponent of a German tradition in geophysics that drew no sharp line between the study of the earth and of its atmosphere.” See: Bowler, Norton History, 399. 25 Draft Plan for the American Geophysical Union - 1987, 21 Oct 1986, Folder 10, Box 27, AGU Records, AIP, College Park, MD. 26 Doel, “American Geophysical Union,” 30–1.

290

system.27 In a statement prepared by AGU president Peter Eagleson as a comment to the US

Congress on the need for federal support and funding for geophysical research, he stated that the

AGU viewed geophysics as, “a truly international science” with interests that include: “Probing

the planets and the sun, the bottom of the ocean and the core of the earth, conducting synoptic

studies of the ocean, the atmosphere and the near space environment[.]”28 Even before the AGU

became quite so panoramic, its members played critical roles in a variety of large-scale Earth

science research programs, including the International Polar Years (1882–3 and 1932–3), and, most prominently, the IGY (1957–8) that brought together over 60,000 scientists from 67 nations for the twentieth century’s largest international scientific project, targeting a wide range of Earth science issues in coordinated fashion. Doel observes that geophysics achieved particular ascendency during World War II and into the Cold War, when the US military and government increasingly linked the discipline to improved understanding of the physical characteristics of strategically important parts of the planet.29 This may help explain why AGU membership grew dramatically throughout the twentieth century, from a mere 75 members in 1922 to over 22,000 by 1989.30 This made it, in 1989, according to AGU President Don Anderson, “the largest U.S.

scientific society devoted entirely to the study of the Earth and its environment in space.”31

In the 1980s, the AGU was a major scientific society. In keeping with geophysics’ legacy of broad-scale research on any and all parts of the Earth and its environs, many of its members worked with new technologies like Earth observing satellites and computer models

27 Draft Plan for the American Geophysical Union - 1987, 21 Oct 1986, Folder 10, Box 27, AGU Records, AIP, College Park, MD. AGU membership included notable planetary scientists like Carl Sagan. See: “Carl Sagan Lecture,” accessed 10 Jun 2019, https://honors.agu.org/sfg-award-lecture/carl-sagan-lecture/. 28 Statement by Peter Eagleson, [2 Feb 1987], Folder 5, Box 12, AGU Records, AIP, College Park, MD. 29 Ronald E. Doel, “Constituting the Postwar Earth Sciences: The Military’s Influence on the Environmental Sciences in the USA after 1945,” Social Studies of Science 35, No. 5 (Oct. 2003): 635–66. 30 Doel, “American Geophysical Union,” 31. 31 Letter from Don L. Anderson to George Bush, 17 Jan 1989, Folder 10, Box 30, AGU Records, AIP, College Park, MD.

291 which were providing new ways to study the planet on a global scale (chapter one). Given these interests in common with the ESSC, the AGU appeared to be a natural “ally.” Indeed, the ESSC frequently targeted AGU members in its promotion of Earth system science, presenting ESS material at AGU conferences and incorporating AGU members into the report review processes

(chapter four).32 However, AGU members responded with lukewarm support for the ESSC’s

ESS research program. While the AGU had a long history of supporting interdisciplinary Earth science research, many members were of the view that the ESSC’s work unnecessarily, even alarmingly, marginalized many areas of geophysics.

Nevertheless, geophysicists were early adopters of the Earth system concept, the phrase promoted by the ESSC as designating an interconnected planet that required interdisciplinary cooperation to promote better understanding. Just as citations for “Earth system” dramatically increased in Science and Nature pre- and post-1985, so too can this trend be observed in AGU publications, including the peer-reviewed Journal of Geophysical Research as well as the publicly-oriented EOS. From 1969 until the end of 1985, these publications contained 31 occurrences of the phrase “earth system.” From 1986 to the present, the phrase appears 10,632 times.33 This is an even more striking trend than those exhibited in Science and Nature publications, indicating the extent to which this phrase was adopted by geophysicists into their research. Clearly, geophysicists could adopt the Earth system concept as a useful description of the planet without fully supporting or agreeing with the ESSC’s proposed ESS program. They could do so because of the vague nature of the Earth system concept, because the Earth system was a boundary object that could be more narrowly defined in local contexts and more generally

32 Telemail from P. Stevens to S. Tilford et al., 3 Jan 1986, Folder 55, Earth System Sciences Collection, NCAR, Boulder, CO; Memorandum from Laura Lee McCauley to [ESSC] Members, Official Agency Liaisons, Technical Liaisons, and Observers, 15 May 1986, Folder 71, Earth System Sciences Collection, NCAR, Boulder, CO. 33 Search conducted on 6 Jun 2019. See: “AGU Journals,” accessed 6 Jun 2019, https://agupubs.onlinelibrary.wiley.com/.

292

defined in interdisciplinary settings. The AGU showed its receptivity to the Earth system

concept despite its ambivalence about ESS in three ways: first, through its promotion of

interdisciplinary Earth science research; second, through the formation of the Earth-as-a-System

Committee (later the Planet Earth Committee); and, lastly, through its attempt to develop its own research program that went beyond ESS to study the “complete” Earth.

A Further AGU Interdisciplinary “Turn”

Though the AGU has a long history of supporting research on all parts of the Earth and its environs, it took work to integrate the kinds of interdisciplinary Earth science that were possible with new tools like satellites and computer models. There was a tension between the newly available tools that elided disciplinary boundaries and traditional practices that compartmentalized Earth science disciplines. The AGU spent considerable resources throughout the 1980s promoting newer cross-disciplinary work. Erik Conway notes that the American scientific community did not consider interdisciplinary research to be “serious science” and, thus, in the 1980s was not in a good position to advance this kind of work. Conway suggests that interdisciplinary work is problematic for scientists because they receive training in a particular specialization with specific techniques—be it atmospheric chemistry, physical oceanography, or ecology—and so they can only provide expertise to a single part of an interdisciplinary research project. This often resulted in early skepticism of this kind of work. No single scientist could be an expert in all parts of an interdisciplinary research project or, perhaps more importantly, in the peer-review process.34

34 Conway notes that interdisciplinary planetary science work received early resistance for this reason. See: Erik M. Conway, Atmospheric Science at NASA: A History (Baltimore, MD: Johns Hopkins University Press, 2008), 242. See also: Ronald E. Doel, Solar System Astronomy in America: Communities, Patronage, and Interdisciplinary Research, 1920–1960 (New York: Cambridge University Press, 1996).

293

Scientists were all-too conscious of these interdisciplinary issues. In January 1986,

Starley Thompson of NCAR communicated to AGU officials his support of interdisciplinary

research focused on interactions between components of the Earth system (see below).

Thompson noted that, “The development and fostering of multi-disciplinary research is always

difficult and often regarded with skepticism by disciplinary specialists.”35 Likewise, climate

scientist Stephen Schneider noted this problem in the first issue of his interdisciplinary Climatic

Change in 1977 (see below). According to Schneider, with the rise of highly specialized science came an essential “arduous disciplinary apprenticeship” that allowed scientists to keep up with new developments and teach the necessary skills to facilitate future developments. However,

Schneider claimed that, in addition to these specialized advances, there was also a need for

“generalists” who had familiarity with more than one field, those who could link research

findings together and communicate them to other specialists.36 In other words, specialization

generated a complementary need for interdisciplinarity.

The AGU was well aware of the issues surrounding interdisciplinarity. A report on the

AGU’s proposed activities for 1987 identified a number of weaknesses the union should address,

which included: “Disincentives to interdisciplinary cooperation within the Union due to

disciplinary loyalties, traditional scientific specialization, and competition for limited research

funds.”37 Interdisciplinary cooperation could not be taken for granted. It had to be continually

promoted within the AGU in order to have any lasting effect. Much of the AGU’s work in the

1980s directly addressed this need. An AGU Planning Committee drafted a three-year plan for

35 Letter from Starley Thompson to Peter Eagleson, 29 Jan 1986, Folder 7, Box 30, AGU Records, AIP, College Park, MD. 36 Stephen H. Schneider, “Climate Change and the World Predicament: A Case Study for Interdisciplinary Research,” Climatic Change 1 (1977): 38-9. 37 Draft Plan for the American Geophysical Union - 1987, 21 Oct 1986, Folder 10, Box 27, AGU Records, AIP, College Park, MD.

294

the Union in October 1984 that outlined program objectives to be achieved by the end of 1987.

One of these organizational objectives included setting, “a strategy for how new areas and

interdisciplinary areas such as polar programs, permafrost/glaciology and geosphere/biosphere

should be handled by the Union.” Achieving this objective might require establishing “new

Sections, committees, affiliated societies or other structural elements” within the AGU.38 A willingness to restructure the AGU organization to advance interdisciplinary aims illustrates the commitment of AGU members to this scholarly approach. In December 1984, the AGU

Executive Committee agreed to hold a half-day follow-up conference at its spring meeting in

May 1985 that would, “focus on strengthening interdisciplinary activities within the Union.”39

One of the earliest and most significant efforts made by the AGU to promote

interdisciplinary research involved the formation of new interdisciplinary journals. These

journals specifically targeted interstitial spaces between, and the connections among, Earth and

environmental science disciplines. In 1980, the AGU’s Publications Committee recommended

that the Union establish two new journals: “Tectonics” and “Geochemical Cycles.” Both of

these journals would be, “problem-area oriented, and are intended to be of strong

multidisciplinary emphasis.” According to the Publications Committee, “these journals are

scientifically worthwhile, are focused on rapidly growing fields which need specialized

publications, and will probably be a great success.”40 The proposed journal on “Geochemical

Cycles” is relevant here, as it was envisioned to provide a venue for those researchers whose

work touched on what today are referred to as biogeochemical cycles, that is, the flows of

chemicals like nitrogen, oxygen, sulphur, and phosphorus through the air, land, organisms, and

38 AGU 3-Year Plan, 29 Oct 1984, Folder 4, Box 12, AGU Records, AIP, College Park, MD. 39 AGU Executive Committee Minutes, 5 Dec 1984, Folder 5, Box 12, AGU Records, AIP, College Park, MD. 40 “New Journals,” [1980], Folder 19, Box 12, AGU Records, AIP, College Park, MD.

295

water. Vladimir Vernadsky and G. Evelyn Hutchison were pioneers in this field (chapter one).

A prospectus on this proposed journal by biogeochemist James Walker observed that, “There has

been increasing interest, in recent years, in the processes that control the overall chemical

composition of the terrestrial environment.” He further noted that this interest was not just

reflected in published research, but also in organizations like NASA and the NSF that had recently held interdisciplinary scientific meetings in this area. Despite this increase in interest, there was, as yet, no single journal in which scientists could publish their research. They were

instead forced to publish in journals dedicated to traditional Earth science disciplines like

geology, geochemistry, oceanography, atmospheric science, and ecology. The new journal

would publish research that dealt with, “processes that influence [the] global chemical

environment, particularly processes that involve the interaction of two or more of the reservoirs:

atmosphere, ocean, biota, and rocks….The emphasis on interaction between reservoirs is

intended to foster an interdisciplinary flavor.” Walker listed some “Illustrative Articles” that

included: “Atmospheric response to deep-sea injections of fossil-fuel carbon dioxide,” “Methane

in marine sediments,” “Ozone fluxes to tobacco and soil under field conditions,” and “Abiotic

and biotic factors in litter decomposition in a semiarid grassland.”41

Tellingly, the AGU went ahead with Tectonics in 1982, while it would take another five

years for the Union to establish a journal emphasizing biogeochemical processes. By this time,

the ESSC’s work was well underway. In 1985, AGU members reopened the discussion about a

journal on biogeochemical cycles after the ICSU announced the formation of the IGBP. At this

point, Jack Eddy (Chairman of the US IGBP Committee and frequent commentator on ESSC

draft reports) wrote to the AGU encouraging support for this endeavour: “I hear more and more

41 Prospectus: New Journal Dealing With Global Chemical Cycles, by James C.G. Walker, Mar 1980, Folder 19, Box 12, AGU Records, AIP, College Park, MD.

296

support for the interdisciplinary treatment of problems of global chemical cycles. It seems to be

an area of science whose time has come, and the best case yet for genuine interdisciplinary

cooperation in front-line science.”42 An updated proposal for a biogeochemical journal in 1986

acknowledged that Eddy had volunteered his editorial services and listed a number of other

possible candidates, including Ralph Cicerone (atmospheric scientist), Michael McElroy

(atmospheric and environmental scientist), Harold Mooney (ecologist), and Roger Revelle

(oceanographer).43 By May 1985, oceanographer (and ESSC draft report reviewer) James

McCarthy had agreed to serve as editor.44 With the final name, Global Biogeochemical Cycles,

the journal was scheduled to begin publication in the fall of 1986 but the first issue did not

appear until March 1987.45 AGU’s Executive Director Fred Spilhaus blamed the delayed

startup, in part, on the fact that there was not yet a “well-established community of authors and

potential readers.” Spilhaus further noted the “close connection” between this journal and the

“new Earth-as-a-System Committee” (see below).46 While the AGU did not endorse the ESSC’s

specific ESS research program, the new journal indicated support for nurturing research that

treated the planet “as a system.”

As mentioned, in the 1980s AGU members identified the “geosphere/biosphere” as an

area ripe for interdisciplinary inquiry.47 This included the development and promotion of

research that incorporated both the Earth and planetary sciences and, for the first time, the life

42 Quoted in: Proposal: AGU Journal Publishing in Global Chemical Cycles and Global Change, [1985], Folder 19, Box 12, AGU Records, AIP, College Park, MD. 43 Proposal: AGU Journal Publishing in Global Chemical Cycles and Global Change, [1985], Folder 19, Box 12, AGU Records, AIP, College Park, MD. 44 Notes: Presidential Report, [May 1985], Folder 4, Box 12, AGU Records, AIP, College Park, MD. 45 Attachment A: Open [AGU] Executive Committee and Council Action Items, 30 May 1986, Folder 4, Box 13, AGU Records, AIP, College Park, MD; “Global Biogeochemical Cycles,” accessed 17 Jun 2019, https://agupubs.onlinelibrary.wiley.com/journal/19449224. 46 Letter from A.F. Spilhaus to James J. McCarthy, 20 Aug 1986, Folder 5, Box 28, AGU Records, AIP, College Park, MD. 47 Draft Plan for the American Geophysical Union - 1987, 21 Oct 1986, Folder 10, Box 27, AGU Records, AIP, College Park, MD.

297

sciences.48 Prior to the 1980s, there were sporadic calls from AGU members to incorporate the

life sciences and ecology into geophysical research concerns, as a necessary component of a

comprehensive understanding of the planet.49 In the 1980s, these concerns became more widespread and formalized as connections were explored between the AGU and the US’s preeminent ecological society, the Ecological Society of America (ESA). Ecology, as the study of the relationships between organisms and their environments, is akin to geophysics in that it has a natural propensity towards interdisciplinarity. Founded in 1915, the ESA’s first members came from 12 different disciplines: plant, animal, and marine ecology; forestry; entomology; agriculture; plant physiology; plant pathology; climatology; geology; animal parasitology; and soil physics.50 In October 1988 Harold Mooney, President of the ESA, wrote to Spilhaus

requesting that the AGU consider, “establishing a more formal relationship” with the ESA.

Mooney supported Spilhaus’ idea to establish a “joint special program committee” that could be

operational by the spring of 1989. At minimum, the committee would organize a joint

symposium at the AGU meeting in fall 1989, though Mooney suggested that this would be

merely the beginning: “Hopefully, the joint committee may have even more ambitious plans.”51

Given the interdisciplinary history of the ESA, this alliance made good sense to Mooney and,

clearly, AGU members agreed. In December 1988, the AGU’s Executive Committee approved

the formation of the joint program committee.52

48 Ibid. 49 In 1967, the Transactions: AGU contained a piece on “Ecology: The Sobering Science.” It argued that ecology deserves, and might even require, increased attention from geophysics since it is concerned with the interconnected relationships between organisms and their environments, which could help geophysicists better understand the interconnected nature of many of the Earth’s processes and increasing environmental concerns. Transactions: AGU article, “Ecology: The Sobering Science,” Sep 1967, Folder 17, Box 82, AGU Records, AIP, College Park, MD. 50 Robert L. Burgess, “The Ecological Society of America: Historical Data and Some Preliminary Analyses,” Ecological Society of America, accessed 11 Jun 2019, https://www.esa.org/history/documents/BurgessHistory.pdf. 51 Letter from Harold A. Mooney to A.F. Spilhaus, 24 Oct 1988, Folder 9, Box 13, AGU Records, AIP, College Park, MD. 52 Actions of the Executive Committee, 4 Dec 1988, Folder 9, Box 13, AGU Records, AIP, College Park, MD.

298

In addition to the ESA, the AGU sought to formalize relations with other major scientific

organizations in the 1980s, notably the American Meteorological Society (AMS). Meteorology

and atmospheric science more generally are closely related to geophysics, which incorporates

physical studies of the atmosphere as part of its expansive reach.53 Though related, the fields

have different historical roots, and therefore practitioners formed different scientific societies to

serve the professional needs of practitioners. But with a number of common concerns, it is also

not surprising that, just as with the ESA, in the 1980s the AGU moved to forge formal relations

with the AMS. Setting an example that was later followed by the AGU in its relationship with the ESA, the AMS Council and the AGU’s Executive Committee approved a plan to hold a joint meeting in 1986.54 By 1988, the AGU Executive Committee—almost like a parent looking for

“playdates”—was considering proposals for “activities” with the AMS that included a “joint

committee on Earth as a System” and a “joint symposium on global change at the 1990 AMS

meeting.”55 Though the AGU and AMS never did form a joint “Earth as a System” committee,

the planned global change symposium closely aligned with the many other global change

conferences, symposiums, and forums held in the late 1980s and early 1990s.56

Not limited to just scientific research activities, the AGU lent its interdisciplinary support

to the PBS documentary series “Planet Earth.” This seven-part television show was produced by

WQED Pittsburgh, in association with the National Academy of Sciences, and began airing on

January 22, 1986. As well as being edifying entertainment, the series was also designed to serve

as part of a “college-credit geoscience telecourse,” covering topics on the Earth’s surface, core,

53 Meteorology is a subset of the atmospheric sciences, with an emphasis on weather forecasting. 54 Letter from Peter Eagleson to Marshall E. Moss and Ralph Cicerone, 9 Oct 1986, Folder 10, Box 27, AGU Records, AIP, College Park, MD. 55 Executive Committee Meeting Preliminary Agenda, 16 Mar 1988, Folder 13, Box 29, AGU Records, AIP, College Park, MD. 56 Folder 15, Box 88, AGU Records, AIP, College Park, MD.

299

mantle, climate, and resources, as well as its relation to other planets and the effects of human

activities.57 Recognizing the “educational importance” of the show, the AGU provided “seed

money” for its production in the early 1980s.58 At the AGU Executive Committee meeting in

December 1985, members recommended that, “a strong effort be made to encourage

congressmen and their staffs to watch this series.”59 With episode titles like The Living

Machine, The Blue Planet, The Climate Puzzle, and The Solar Sea, the “Planet Earth” series promoted a view of the planet in line with the AGU view of Earth as a requiring comprehensive study from the its core to its space environs (see below). The show brought in

AGU members like geologist Brent Dalrymple and seismologist Don Anderson as well as James

McCarthy, eventual editor for the Global Biogeochemical Cycles journal. A review of the

textbook accompanying “Planet Earth” written by ESSC and AGU member Kevin Burke noted

that the AGU was “heavily committed” to this television show.60

The Chapman Conference on Gaia

This sketch of the AGU’s interdisciplinary activities in the 1980s shows that it was, like the

ESSC, orienting Earth scientists towards interdisciplinary studies of the planet as an

interconnected system, though there was no generally accepted way to talk about the Earth “as a

system.” The Chapman Conferences provide further evidence that the AGU was working in this

semantic void and was, therefore, particularly amenable to the ESSC’s Earth system concept.

Every spring and fall, the AGU holds annual meetings. They are large affairs, with thousands

57 Lee Margulies, “TV Review: PBS ‘Earth’ Series Off to an Earthshaking Start,” Los Angeles Times, 22 Jan 1986, accessed 23 Jun 2019, https://www.latimes.com/archives/la-xpm-1986-01-22-ca-31688-story.html. 58 “More About ‘Planet Earth,’” EOS 67, no. 4 (28 Jan. 1986): 43. 59 [AGU] Executive Committee Minutes, 11 Dec 1985, Folder 8, Box 10, AGU Records, AIP, College Park, MD. 60 Kevin Burke, “Planet Earth and the New Geoscience,” EOS 67, no. 4 (28 Jan. 1986): 42-3.

300

attending lectures and visiting display booths.61 The AGU also sponsors smaller conferences

focused on specific aspects of the physical sciences. Of these, the most prominent are the

Chapman Conferences that are held multiple times a year and have been since 1976. Inspiration

for the conference came from the work of Sydney Chapman, an early pioneer in atmospheric and

solar-terrestrial physics and one of the main proponents of the IGY.62 In its guidelines for

Chapman Conference convenors, the AGU reflected that Chapman made, “significant

discoveries in many different disciplines, including earth and atmospheric science; ionospheric,

magnetospheric, interplanetary, and solar physics; and more.” Chapman applied, “basic

principles freely and successfully across traditional boundary lines[.]”63

These conferences were intended to be smaller in scope than the annual AGU meetings,

with just 60 to 100 attendees, not thousands. Chapman Conferences were intended to promote

“newly emerging research fields or problem areas,” and bring together “participants who are

active researchers in diverse but related fields[.]”64 Though these conferences need not be interdisciplinary in scope, the forum was certainly conducive to this. Topics over the years varied widely, coinciding with the broad interests of the AGU, from the center of the Earth to outer space, crossing and interlinking different Earth and interplanetary science disciplines. By the mid-1980s, as many as five Chapman Conferences were held per year. In 1986, topics ranged from “Ionospheric Plasmas in the Magnetosphere” to “Modeling of Rainfall Fields” to

61 In 1988 the AGU annual spring meeting had over 2,700 registrants. The fall meeting was expected to be ten percent higher than the 1986 record, with over 3,800 abstracts received. “AGU 1988,” [1988], Folder 9, Box 13, AGU Records, AIP, College Park, MD. Currently, the fall annual meeting attendance figure is over 23,000. See: “Meeting,” AGU, accessed 12 Jun 2019, https://meetings.agu.org/. 62 Gregory A. Good, “Sydney Chapman: Dynamo Behind the International Geophysical Year,” in Globalizing Polar Science: Reconsidering the International Polar and Geophysical Years, eds. Roger D. Launius, et al. (New York: Palgrave Macmillan, 2010), 177-203. 63 “Chapman Conferences: Guideline for Convenors,” EOS 56, no. 11 (Nov. 1975): 836. 64 Ibid.

301

“El Niño.”65 As early as October 1985, the AGU Planning Committee began arrangements for,

“the first Biological-Geophysical Chapman Conference[.]”66 No Chapman Conference was

organized with this exact name. However, that the AGU Planning Committee tentatively

discussed it as a possibility indicates its interest in bringing together scientists whose research

crossed the boundaries between these two areas. This interest is underscored by the AGU’s later

move to establish more formal ties with the ESA. Other Chapman Conferences focused on

“biological-geophysical” topics even if they did not use that exact phrase.

Perhaps the most famous—or infamous, depending on one’s stance—of these

“biological-geophysical” Chapman Conferences was focused on James Lovelock’s Gaia hypothesis. Since its introduction, Gaia has been more or less controversial, depending on how it has been interpreted and who is doing the interpreting. If the hypothesis simply means that biota contribute to the physical and chemical conditions on Earth, then the scientific community in the

1980s already viewed this hypothesis as well-supported by evidence. It was certainly in line with the growing understanding of the Earth as a system of interconnected components, including living components. At the other extreme, however, was an interpretation of the Gaia hypothesis in which organisms intentionally altered environmental conditions to make them optimal for life. This interpretation was more controversial, to say the least.67 Further, different scientific communities reacted differently to the Gaia hypothesis. Michael Ruse observes that biologists were particularly hostile to Gaia (chapter one).68 Other scientific communities, like

65 The most Chapman Conferences held in a single year was six in 1978. See: “AGU Chapman Conferences,” EOS 68, no. 17 (28 Apr. 1987): 489. 66 (Tentative) Agenda: Planning Committee, Oct. 4-5, 1985, 29 May 1985, Folder 20, Box 12, AGU Records, AIP, College Park, MD. 67 James W. Kirchner, “The Gaia Hypotheses: Are They Testable? Are They Useful?” in Scientists on Gaia, eds. Stephen H. Schneider, Penelope J. Boston (Cambridge, MA: MIT Press, 1991), 38-46. 68 Michael Ruse, The Gaia Hypothesis: Science on a Pagan Planet (Chicago: University of Chicago Press, 2013), 214-5.

302 those in the physical sciences, were more receptive to the hypothesis, even if these scientists usually rejected its most extreme interpretations.

The greater willingness to engage with the discipline-transcending Gaia hypothesis among non-biological scientific communities is suggested by several key events in the 1980s.

First, NCAR hosted James Lovelock as part of its twenty-fifth anniversary celebrations in 1985

(chapter one). Second, the AGU devoted one of its Chapman Conferences to the “Gaia

Hypothesis,” held in San Diego from March 7 to 11, 1988. The major impetus behind this conference came from climate scientist Stephen Schneider, who worked at NCAR throughout the

1980s. Schneider had a storied scientific career from the early 1970s up to his untimely death in

2010. With training in mechanical engineering and plasma physics, Schneider’s postdoctoral research interests in climate modeling took him to NASA’s Goddard Institute of Space Science and then NCAR, where he remained until he joined Stanford’s Department of Biological

Sciences in 1992. At NCAR, Schneider was credited by staff as having played a major role, “in

NCAR’s approach to climate research in the 1970s and 1980s,” eventually heading the center’s

Interdisciplinary Climate Systems Section.69 In 1977, Schneider founded the interdisciplinary journal Climatic Change to serve as, “a means of exchange among researchers from a variety of disciplines who are working on problems related to climatic variations.”70 Up to his death,

Schneider also worked as an author on the International Panel on Climate Change (IPCC) reports. Beyond his scientific research and professorial duties, Schneider engaged in climate change public outreach that included writing the popular Global Warming: Are We Entering the

Greenhouse Century? (1989). For work like this, one colleague claimed that Schneider, “did for

69 “Stephen Schneider: An Extraordinary Life,” NCAR and UCAR News, accessed 13 Jun 2019, https://news.ucar.edu/2270/stephen-schneider-extraordinary-life. 70 Stephen H. Schneider, “Editorial for the First Issue of Climatic Change,” Climatic Change 1, no. 1 (Mar. 1977): 3.

303 climate science what Carl Sagan did for astronomy.”71 Schneider was also well known to ESSC members. He was listed as one of the hundreds of “contributors” to the ESSC’s work,72 and even received an honourable mention in John Imbrie’s “dynos” joke (chapter three).73

Schneider’s background amply points to his interest in interdisciplinary research and global scientific issues. It is therefore not surprising that Schneider proposed a Chapman

Conference on the Gaia hypothesis with the AGU. According to Schneider, the motivation for the proposal came from being involved with a NOVA film crew making a documentary about the Gaia hypothesis in 1986.74 NOVA personnel filmed alternating clips of Schneider and

Lovelock debating various facets of the hypothesis, until Schneider one day declared in exasperation: “this is outrageous. Why am I debating Jim Lovelock via BBC and ‘Nova’ producers? Why don’t we have a scientific meeting? So I proposed to the AGU that we have a

Chapman conference on the Gaia hypothesis.” Schneider expressed skepticism about the hypothesis, but believed that it merited scientific attention: “It was largely ignored or disdained by the scientific community. And I thought that was wrong, because I thought it was a brilliant and clever insight, but probably not right, but certainly worthy of discussion.”75 The Chapman

Conferences were, according to Schneider, the “most prestigious scientific meetings that you could get at the AGU” and so a fitting venue for a discussion. Schneider recalled encountering

71 This quotation is from Benjamin Santer of the Lawrence Livermore National Laboratory. See: “Stephen Schneider: An Extraordinary Life,” NCAR and UCAR News, accessed 13 Jun 2019, https://news.ucar.edu/2270/stephen-schneider-extraordinary-life. 72 ESSC, Closer View, 203. 73 Not only did John Imbrie’s dynapeople have dramatically faster rates of metabolism, they also spoke “21,932 times faster than we do -- even faster than Steve Schneider.” See: Memorandum from John Imbrie to John A. Dutton, 11 October 1984, Earth Observing System Series, Box 18042, ESSC/Bretherton Committee File, NASA Historical Reference Collection, Washington, DC. 74 NOVA is a PBS science television series. According to Schneider, NOVA was making a Gaia documentary after BBC’s “Horizon” had already done one. See: Stephen H. Schneider, interviewed by Robert M. Chervin, 10-13 Jan 2002, Palo Alto, CA, accessed 14 Jun 2019, http://n2t.net/ark:/85065/d72n50p5. 75 For instance, Schneider found the notion of “planetary self-control” and conditions being “good for the biota” ill- defined and potentially conflicting. See: Schneider, interview, 10-13 Jan 2002.

304 resistance from AGU members, including Wallace Smith Broecker. Though Broecker specialized in chemical oceanography and the carbon cycle—and so might have been presumed to be amenable to debating the scientific merits of the Gaia hypothesis—Schneider reported that

Broecker was “outraged” that the AGU might hold such a conference. According to Schneider,

Broecker objected that, “‘Gaia isn’t science! This is ridiculous!’” It took “about a year and a half,” but with “sufficiently persuasive” support from AGU members like Ralph Cicerone and

Juan Roederer, the AGU council finally approved the meeting.76

AGU documents reveal a different account of the lead up to the Chapman Conference on the Gaia hypothesis. They show that the AGU approved holding the conference in May 1986, to be convened by Schneider and atmospheric physicist Glenn Shaw “probably in the fall of 1987.”

However, the AGU Council made sure to add the caveat that, “approval of the Gaia conference does not imply approval or even endorsement of the Gaia hypothesis.”77 So, approval for the meeting was quick, even if the AGU did not endorse the hypothesis. What appears to have taken more time was securing funding and a venue.78 In November 1986, the AGU Committee noted that the Gaia Chapman Conference date and location had not yet been fixed, though Penelope

Boston was now a co-convenor.79 Schneider claimed that he included Boston because he

“wanted to have a biologist” involved in the conference.80 By April 1987, AGU staff reported that the conference was scheduled for March 1988 in San Diego.81 Perhaps part of the delay stemmed from funding issues. A report on grants and contracts for AGU program activities from

76 Schneider, interview, 10-13 Jan 2002. 77 Minutes of the [AGU] Council Meeting, 20 May 1986, Folder 7, Box 10, AGU Records, AIP, College Park, MD. 78 The initial proposed location was Kona, Hawaii, but the AGU’s Executive Committee raised concerns with the expense and travel time required with that venue. See: Minutes of the [AGU] Executive Committee, 19 May 1986, Folder 4, Box 13, AGU Records, AIP, College Park, MD. 79 [AGU] Meetings, Nov 1986, Folder 10, Box 10, AGU Records, AIP, College Park, MD. 80 Schneider, interview, 10-13 Jan 2002. 81 “Chapman Conferences,” Apr 1987, Folder 12, Box 10, AGU Records, AIP, College Park, MD.

305

May 1988 listed four grant requests for the Gaia Chapman Conference made to the NSF, the

MITRE Corporation, the USGS, and NASA. Only the NSF and MITRE provided funds, with

NASA and the USGS declining. All the other Chapman Conference grant requests listed in this

report received funding from external organizations, conferences on seismic anisotropy, the fate

of particulate and dissolved components within the Amazon dispersal system, plasmic waves and

instabilities, perovskite, ion acceleration in the magnetosphere and ionosphere, magnetotail

physics, fast glacier flow, and El Niño. Only grant requests for the Gaia conference were

declined by external organizations, indicative of the scientifically controversial character of

Gaia.82

What is interesting and important about the AGU’s Chapman Conference on the Gaia hypothesis is that Schneider explicitly viewed “Gaia” as a phrase that could potentially fill the semantic void in the Earth sciences. Schneider recognized the need for an integrating concept that could describe the Earth as an interconnected system and motivate future interdisciplinary research. For Schneider, the Gaia hypothesis was not interesting because it was true. In fact, it

lacked many details that might have rendered it more scientifically rigorous and, as noted above,

Schneider thought that the hypothesis would probably turn out to be false. But it might provide

him with a turn of phrase that could direct the interdisciplinary efforts of Earth scientists towards

studying the Earth as a system. Schneider believed that “Gaia” might be useful as a heuristic

device. It could, “advance science because people are passionate about the ideas. And when

they are passionate they are going to go out and study, and they are going to look at the earth the

right way—as a system, not as a disconnected set of disciplines.”83

82 Report on Grants and Contracts in Conduct Program Activities, 1 May 1988, Folder 7, Box 13, AGU Records, AIP, College Park, MD. 83 Schneider, interview, 10-13 Jan 2002 [emphasis added].

306

For Schneider, the Gaia hypothesis might align scientific concerns under the banner of studying the Earth in the “right way,” which meant studying the Earth “as a system.” As it would turn out, the hypothesis was simply too controversial, particularly among biologists to ever fill this role. The “Earth system” concept was able to better fill the void than “Gaia” since it came with much less vitalist and teleological baggage. Its very strength came from its innocuously vague nature. As a boundary object, the Earth system could broadly refer to the

Earth as an interconnected planet, while at the same time it could be defined in more specific ways in disciplinary contexts. Vagueness here facilitated interdisciplinary communication and specialist use, without any particular interpretation causing undue controversy in the Earth science community at large. Though a geophysicist and an atmospheric scientist might differently define the Earth system and place its boundaries in different locations, they both recognized that the reasons for these differences resulted from differing scientific interests, and not from anything inherently unscientific in any of these definitions.

The Gaia hypothesis, on the other hand, was problematic because its extreme interpretations were quite controversial, to the point of being viewed by many as simply unscientific.84 One Gaia Chapman Conference presenter, philosopher and physicist James

Kirchner, identified at least five different interpretations of the Gaia hypothesis. They ranged from the broadly accepted “soft” Gaia—that merely suggested that biota can alter certain conditions in abiotic environments—to the much more controversial “hard” Gaia—that argued that biota modified the abiotic environment to produce optimal conditions for life. Much of the scientific hostility to the Gaia hypothesis was a reaction to the more extreme, or “hard,”

84 Ronald Kline describes a similar problem for the field of cybernetics in the 1960s, when it was taken up by those promoting pseudoscientific ideas like scientology. This tarnished the field, and helps explain why, for Kline, we today live in an “Information Age” rather than a “Cybernetic Age.” See: Ronald R. Kline, The Cybernetics Moment: Or Why We Call Our Age the Information Age (Baltimore, MD: Johns Hopkins University Press, 2015), ch. 3.

307

interpretations.85 Writing in EOS, Eric Kauffman stated that “soft” Gaia interpretations “fared well….But then these are old hypotheses, restated in the context of Gaia, that reflect the ecological structure of life systems and evolutionary adaptation to the physical and chemical constraints of Earth systems.” The “harder” interpretation, in contrast, “will still generate fierce debate for years to come.”86

If “Gaia” had not proved so problematic, Schneider might have been right and it might

have been the phrase that filled the semantic void in the Earth sciences. Today, scientists might

have used “Gaia” rather than the “Earth system” when referring to the planet as a scientific

object. In a more recent publication on Gaia, Lovelock uses “Gaia” and “Earth system”

interchangeably, as if Gaia was simply another term for the Earth system, simply another way to

broadly refer to an interconnected planet.87 However, Gaia and the Earth system are not merely

titularly different. There are important distinctions. Whatever Lovelock might claim, other

scientists take Gaia to be much more problematic. Gaia’s “hard” interpretation—that ascribes

purposeful regulatory action and optimization to organisms and planetary conditions—is beyond

the scientific “pale” for many scientists. The Earth system concept did not have these

problematic connotations. Admittedly, both Gaia and the Earth system are vague concepts that

have multiple interpretations. But vagueness for the Gaia hypothesis, unlike for the Earth system

concept, was not a strength. Gaia’s vagueness meant that it would always, for some, be linked to

“hard” interpretations with questionable affiliations with vitalism and teleology. It is

conceivable that Gaia could have filled the semantic void in the Earth sciences, but its

85 Kirchner, “The Gaia Hypotheses,” 38-46. 86 Eric G. Kauffman, “The Gaia Controversy: AGU’s Chapman Conference,” EOS 69, no. 31 (2 Aug. 1988): 764. 87 See: James Lovelock, The Vanishing Face of Gaia: A Final Warning (Toronto: Penguin Books, 2010).

308 contentious features prevented it from actually doing so. Instead, it was the innocuously vague and thereby quite useful Earth system concept that filled the void.

Earth-as-a-System

The AGU’s active support for interdisciplinary research in the Earth sciences throughout the

1980s helps explain its receptivity to the ESSC’s Earth system concept that described an interconnected Earth requiring interdisciplinary research. AGU members supported research that did not fit neatly into a single discipline, that crossed boundaries, and that made connections between research and practitioners previously separated by the need for specialization and a lack of tools to facilitate collaborations. With more comprehensive satellite datasets and the computational capacity to calculate ever more interactions between the air, oceans, land, and biota, interdisciplinary Earth science research was now possible in the 1980s in a way it had not been before. The AGU’s collaboration with other national science organizations like the AMS and the ESA, along with its interdisciplinary conferences like the Gaia Chapman Conference, suggest that members viewed the various Earth science disciplines as researching different components of what was, ultimately, a single Earth system. Given this, it is not surprising that

AGU members were particularly susceptible to the ESSC’s extensive promotional activities and, therefore, quick to adopt the ESSC’s phrase “Earth system” to denote an interconnected planet that could now be studied scientifically. The phrase was usefully vague in a way that Gaia was not. There were no scientifically problematic ways to interpret the Earth system. What the Earth system was simply depended on research interests, and in particular the timescales of the processes being studied. That the AGU was quick to adopt the Earth system concept while simultaneously finding the ESSC’s ESS research program problematic is best illustrated by the special committees formed within the AGU in the latter part of the 1980s, first the Earth-as-a-

309

System Committee and later the Planet Earth Committee. Both committees were formed to

promote and support research on what the AGU called the “whole” Earth system, partly in

reaction to the ESSC’s focus on timescales of decades to centuries (chapter three).

As early as 1984, AGU members began discussions about the promotion of research that studied the “Earth as a planet.” AGU member and upper atmospheric physicist Juan Roederer first proposed the idea at a December 1984 AGU leadership conference.88 This idea—to study

the Earth “as a planet”—emerged out of the high profile planetary science missions that flew in

the 1960s and 1970s.89 Planetary flybys, orbiters, and landers of the Mariner and Viking

Programs collected a wide variety of physical and chemical data on Mercury, Mars, and Venus.

Unlike many of the early Earth observing satellites that flew with a limited number of

instruments targeting specific physical, chemical, or biological variables, planetary mission

platforms often carried more instruments to collect a breadth of planetary data. For instance,

Landsat 1 (launched in 1972) carried only two sensors: a return beam videocon and a

multispectral scanner that collected data in four spectral bands.90 The Viking 1 lander, which

arrived at Mars in 1976, carried seven scientific instruments to collect seismological, biological,

chemical, meteorological, and other geophysical data.91 Some claim that these planetary

missions motivated more holistic studies of the Earth.92 In one account, in 1976, shortly after the

landing of Viking 1, leading Mars specialists convened for a three week meeting. After

88 Letter from Sean C. Solomon to A.F. Spilhaus, 1 Apr 1985, Folder 5, Box 12, AGU Records, AIP, College Park, MD. 89 The origins of the phrase “Earth as a planet” trace back at least to a series of books published on “The Solar System.” Volume II was titled “The Earth as a Planet” and contained essays by scientists on, “those aspects of geophysics, geochemistry, and atmospheric physics as pertain to the earth as a whole[.]” See: Gerard P. Kuiper, “Preface,” in The Earth as a Planet, ed. Gerard P. Kuiper (Chicago: University of Chicago Press, 1954), v. 90 “Landsat 1,” NASA Landsat Science, accessed 18 Jun 2019, https://landsat.gsfc.nasa.gov/landsat-1/. 91 Asif A. Siddiqi, Beyond Earth: A Chronicle of Deep Space Exploration, 1958-2016 (Washington, DC: NASA, 2018), 129. 92 Erik Conway directly links studies of other planets to climate science on Earth. See: Erik M. Conway, “Planetary Science and the ‘Discovery’ of Global Warming,” in Exploring the Solar System: The History and Science of Planetary Exploration, ed. Roger D. Launius (New York: Palgrave Macmillan, 2013), 183-202.

310 discussing new experiments and observations for future missions, atmospheric chemist Michael

McElroy reportedly suggested something that he had been telling scientists for many years: “You know, we’ve never done anything like this for the earth.”93 According to McElroy, scientists had never conducted a systematic study of the planet Earth—they had never treated it as a planet—in the same way they had treated other planets in the solar system. To study the Earth “as a planet” implied another “Copernican turn.” That is, the Earth needed to be studied in the same way as other planets, understanding its present state or condition as the outcome of interconnected processes still in effect. This implied that satellites might look at Earth just as they look at Mars and Venus to collect data on the conditions that define “.” In the specific case of the Earth, this would be to define the “Earth system.”

The AGU took up Roederer’s recommendation to study the Earth “as a planet” in 1985.

Sean Solomon—geophysicist and primary author of the “Solomon” report (chapter three)94— wrote to Spilhaus in April 1985 regarding Spilhaus’ “request for interdisciplinary topics” to be discussed at the AGU Council meeting in May. Solomon recommended discussion of the “Earth as a planet” idea that “generated enthusiastic response” from AGU members in December.

According to Solomon, “The topic crosses all section lines, and the most important advances are likely to come from groups and individuals with broader overview and expertise.” Studying the

“Earth as a planet” would, according to Solomon, join a number of other similar initiatives, including “NASA’s ESSC activities...and it is in the interest of AGU to formulate a position on these initiatives and to help to see that the best ideas are implemented.”95 This suggests that, while Solomon was supportive of the ESSC’s “activities,” the AGU should form its own group

93 Quoted in: Edward Edelson, “Laying the Foundation,” Mosaic 19, no. 3-4 (1988): 6. 94 Committee on Earth Sciences of the Space Science Board, National Research Council, A Strategy for Earth Science from Space in the 1980s, Part I: Solid Earth and Oceans (National Academy Press, Washington. DC, 1982). 95 Letter from Sean C. Solomon to A.F. Spilhaus, 1 Apr 1985, Folder 5, Box 12, AGU Records, AIP, College Park, MD.

311

to make sure that the concerns of geophysicists received adequate attention. Spilhaus clearly

agreed with Solomon. At the AGU’s Executive Committee meeting in May 1985, it was decided

that, in October, the AGU’s Planning Committee would consider whether the AGU should form

a “Union-wide committee in [the] area of ‘Earth-as-a-system’” and plan for a series of EOS articles on “Earth-as-a-system.”96

This word choice for the committee name closely resembles the language used by the

ESSC. The May decision to hold discussions in October was made at the same annual meeting where the ESSC organized an entire panel on ESS. Called “A Mission to Planet Earth,” the ESS panel was chaired by atmospheric scientist and ESSC member Ronald Prinn. It included six presentations by other ESSC members on components of the Earth system.97 Perhaps playing to the geophysical audience, the choice of name for this panel also drew on an opinion piece that

NASA’s Associate Administrator for the OSSA, Burton Edelson, wrote for Science in January

1985. Edelson wrote the piece after attending ESSC meetings that introduced him to the group’s ideas for a large-scale Earth science program.98 Drawing on the planetary story mentioned above, Edelson claimed that, “In some ways we know more about neighbouring planets than we do about the earth….Of course, through the centuries we have accumulated a mountain of detailed data points and much phenomenological knowledge about the earth and the constituents of its geosphere and biosphere. However, we lack synoptic, systematic, and temporal knowledge of our own planet and an understanding of the mechanisms underlying the global processes that affect it.” Clearly drawing on ESSC language, Edelson called for a “mission to planet Earth”

that made use of new technologies to “study the earth as a system” in order to “gain

96 (Tentative) Agenda, [AGU] Planning Committee, Oct. 4-5, 1985, 29 May 1985, Folder 20, Box 12, AGU Records, AIP, College Park, MD. 97 AGU Presentation: May 29, 1985, n.d., Folder 136, Earth System Sciences Collection, NCAR, Boulder, CO. 98 Participant List as of 8 June 1984, Fourth Meeting, ESSC Summer Study, Charlottesville, Virginia, 8 Jun 1984, Folder 28, Earth System Sciences Collection, NCAR, Boulder, CO.

312 comprehensive knowledge, not only of the state of the earth system and of global processes, but also of changes in state and processes.”99

With this influence from the ESSC and its ideas, in December 1985 the AGU Council approved the formation of a coordinating committee on “Earth-as-a-System.” It formed part of what the AGU considered its “accomplishments” in the area of “interdisciplinary science.”100 A number of individuals wrote to AGU President Peter Eagleson expressing interest in joining the new committee.101 One prospective member—Valery Lee, a physical oceanographer, formerly at NOAA and by 1986 a data program manager at the NSF—directly linked the Earth-as-a-

System Committee to the ESSC. In her letter expressing her interest, Lee cited her experience with the ESSC as a major qualification for membership: “during my two-year tenure as ‘satellite policy analyst’ at NOAA headquarters, I worked closely with Francis Bretherton’s NASA advisory committee on Earth System Science. In this role I was party to, and participant in, the many debates on what is earth system science and what we should be doing about it.”102 Indeed,

Lee was actually a member of the ESSC “Writing and Support Group” that formed in the latter part of 1985 to prepare what would become the ESSC’s Overview report.103 Though Lee was

99 Burton I. Edelson, “Mission to Planet Earth,” Science 227 (25 Jan. 1985): 367. “Mission to Planet Earth” would later be the name NASA used to describe its planned program to study the Earth comprehensively via satellites, in- situ studies, data analysis, and modeling. It included the massively expensive Earth Observing System (EOS). In 1989, it received Congressional approval and appropriations, becoming the most expensive science program ever in the US. However, the name was rather short-lived, since funding and science issues in the 1990s led to numerous redesigns, rescopings, rebaselinings, and restructurings of the mission, along with much political acrimony regarding the necessity and largess of the program. According to Eric Goldstein, in 1998, in an effort to “cloak the program in a new bipartisan spirit,” NASA changed the name of the program to Earth Science Enterprise. See: Conway, Atmospheric Science at NASA, 199, 243; Goldstein, “NASA’s Earth Science Program,” 218. 100 [AGU] Council Meeting, 17 May 1988, Folder 13, Box 10, AGU Records, AIP, College Park, MD. 101 Folder 7, Box 30, AGU Records, AIP, College Park, MD. 102 Letter from Valery E. Lee to Peter S. Eagleson, 25 Feb 1986, Folder 7, Box 30, AGU Records, AIP, College Park, MD. 103 This group also included: Ray Arnold, James Baker, Paul Blanchard, Francis Bretherton, John Dutton, Laura Lee McCauley, Berrien Moore, Stan Ruttenberg, Shelby Tanner, Shelby Tilford, and Kathy Wolfe. See: ESSC Writing and Support Group: August 1985, 1 Aug 1985, Folder 10, Earth System Sciences Collection, NCAR, Boulder, CO.

313

not selected, of the final 12 members for the Earth-as-a-System Committee, two—James

McCarthy and Ronald Prinn—were also ESSC members.

Though the AGU may have had grand designs for the Earth-as-a-System Committee that went beyond the narrower focus of the ESSC, it did not live up to expectations. Unlike the

ESSC, the committee never developed a clearly defined goal that was coherently and uniformly understood by committee members or the AGU more broadly. The original proposal noted the need to study Earth in a way that transcended traditional disciplinary considerations by focusing on interactions between “the earth sciences, ocean sciences, atmospheric science, solar-terrestrial physics and biology” on global scales. Influenced by a series of briefings on AGU interdisciplinary activities at the fall 1985 meeting—including the ESSC’s session on “A

Mission to Planet Earth”—the proposal stated that scientists needed, “to build an understanding of the earth as an interactive system as a basis for useful and timely forecasts of future change.”

Members believed that the AGU, “has much to contribute to and gain from these efforts, and that the creation of an interdisciplinary committee tasked with pursuing the opportunities was in order.” The Earth-as-a-System Committee would be charged with developing, “research focused on earth as a system including aspects of that are most directly coupled to geophysics.” In contrast to the ESSC’s proposal for ESS, geophysical concerns would be a central, rather than peripheral, component of this research program. Its responsibilities would include defining appropriate problems and research tasks for AGU activity, making appropriate connections with scientific organizations specializing in biology, educating AGU members on relevant issues, and communicating with funding agencies.104 But what science, exactly, should

be considered “appropriate” for the committee to promote? How should the committee properly

104 Draft Proposal for Establishment of an AGU Inter-Sectional Coordinating Committee on Earth-as-a-System, [1986], Folder 6, Box 30, AGU Records, AIP, College Park, MD.

314

engage funding agencies? What was the best way to educate AGU members on these issues?

What were the “Earth-as-a-System” issues most in need of communication? Unfortunately for

the AGU, the committee failed to provide any adequate answers.

Early on, the committee appears to have done little substantive work, apart from holding

meetings during the AGU’s fall and spring meetings in 1986 and 1987.105 One member,

oceanographer Donald Olson, wrote to Eagleson in April of 1988 to register his concern, “over

the future focus of the [Earth-as-a-System] committee[.]” Olson commended Committee Chair

Raymond Price’s efforts, but suggested that the committee suffered from a “lack of tangible

goals and duties.”106 Eagleson responded: “I share your concern and feel that you may be right

about the need for a more specific charge.” He further noted that the committee was being

“reconstituted” by Don Anderson.107 As early as December 1987, the AGU Executive

Committee agreed to “re-evaluate the Earth-as-a-System Committee charge and seek a new chairman[.]”108 This reconstitution transformed the committee into the Planet Earth Committee

(PEC), shedding the original name and adding more specificity to the committee’s mandate. The

primary goals of the PEC would be to determine: “what to measure and how; how to collect the

data; what are the priorities; etc. The Committee will also inform governmental decision makers

and AGU members of its actions and will seek ways to gain implementation of the Initiative.”109

105 Preliminary Agenda, AGU Leadership Conference, 18 May 1986, Folder 2, Box 28, AGU Records, AIP, College Park, MD; Calendar of Events (other than scientific sessions), 1986 Fall Meeting -- San Francisco, CA, December 7- 10, [1986], Folder 4, Box 13, AGU Records, AIP, College Park, MD; AGU Executive Committee Agenda, 18 May 1987, Folder 3, Box 13, AGU Records, AIP, College Park, MD. 106 Letter from Donald B. Olson to Peter S. Eagleson, 12 Apr 1988, Folder 8, Box 13, AGU Records, AIP, College Park, MD. 107 Letter from Peter S. Eagleson to Donald B. Olson, 20 Apr 1988, Folder 8, Box 13, AGU Records, AIP, College Park, MD. 108 [AGU] Executive Committee Action Items, 9 Dec 1987, Folder 6, Box 13, AGU Records, AIP, College Park, MD. 109 [AGU] Executive Committee Minutes, 15 May 1988, Folder 14, Box 10, AGU Records, AIP, College Park, MD.

315

In large part, the Earth-as-a-System Committee was overhauled because, in the unflattering words of AGU official Les Meredith, the committee “accomplished almost nothing[.]”110 However, another important reason for the reconstitution rested in the name of the committee itself. The original “Earth-as-a-System” nomenclature is quite similar to that used by

NASA’s Earth System Sciences Committee. As early as May 1986, the AGU Executive

Committee agreed to write to chairman Price suggesting that the name of the Earth-as-a-System

Committee, “may need reconsideration as it is rather close to NASA’s Earth Systems Science activity.”111 This suggestion was raised again by Eagleson and agreed upon at an AGU

Executive Committee meeting in September 1986.112 The AGU felt pressure to distinguish its

Earth-as-a-System Committee from NASA’s ESSC arguably because it defined the Earth system

differently. While the AGU agreed that it was important to study the Earth as a system, and it

adopted the Earth system phrase as a way to describe the planet, it did not support the ESSC’s

specific ESS research program that focused predominantly on shorter term processes occurring

in the fluid portions of the Earth. Spilhaus wrote to Earth-as-a-System Committee Chair Price in

November 1986 to stress the importance of transcending the ESSC’s research program: “I sense that one of the important issues is how can AGU make clear that ‘Earth System Science’ activity must include the whole earth and not just that portion that interacts on a day-to-day basis with the fluid earth. A clear statement with scientific rationale may be necessary to supplement the

110 Telemail from [Les Meredith] to Brent [Dalrymple], [30 Jan 1989], Folder 12, Box 20, AGU Records, AIP, College Park, MD. Les Meredith was the AGU’s Group Director, Research Programs and Meeting and Member Programs. See: AGU Planet Earth Committee Roster, July 1, 1989 - June 30, 1990, Folder 9, Box 20, AGU Records, AIP, College Park, MD. 111 AGU Executive Committee Meeting Minutes, 22-23 Sep 1986, Folder 10, Box 10, AGU Records, AIP, College Park, MD. 112 [AGU] Executive Committee Actions - September 22-23, 1986, 27 Oct 1986, Folder 10, Box 27, AGU Records, AIP, College Park, MD.

316

Bretherton [ESSC] report.”113 For AGU members, the ESSC research program contained a

tragic flaw: its Earth system science did not go far enough. It was too narrowly focused on the

“fluid” portions of the Earth, to the neglect of solid Earth processes. While the ESSC had

attempted to compromise on this issue by organizing research around two wiring diagrams of the

Earth system for long- and short-term timescales—thereby instantiating two different “Earth

systems”—clearly this compromise did not go far enough for the AGU.

Planet Earth Committee

Prior to the transformation of the Earth-as-a-System Committee into the Planet Earth Committee, the AGU began a parallel activity called the Planet Earth Initiative (PEI) in early 1988. Similar to the Earth-as-a-System Committee, PEI achieved very little. However, the broad descriptions of the proposed PEI initiative demonstrate the AGU’s ambition to eclipse the ESSC’s emphasis on shorter term processes via an international research program to study the “Earth as a total system.”114 “Total” meant studying the processes of the Earth system that occurred on all

timescales. Chaired by atmospheric scientist Ralph Cicerone, a group of AGU members

prepared a “PEI Statement” that was approved by the AGU council in April 1988. It was further

developed to include a series of recommendations for politicians and government agencies.

These recommendations revolved around the AGU’s suggestion that the US take “the initiative”

in building a global observing system and active research program that, “will allow scientists to

understand the natural and anthropogenic changes in the Earth’s environment and the global-

113 Letter from A.F. Spilhaus to Raymond A. Price, 12 Nov 1986, Folder 10, Box 27, AGU Records, AIP, College Park, MD [emphasis added]. 114 Planet Earth Initiative, [1987], Folder 11, Box 20, AGU Records, AIP, College Park, MD [emphasis added].

317

scale forces that control them.”115 To achieve these objectives, the PEI statement recommended,

“a global scientific program aimed at obtaining a fundamental understanding of Planet Earth.”

While there had always been local and regional knowledge of the planet, “Recently we have

begun to recognize that global-scale forces control earthquakes, volcanoes, weather and climate,

and that to manage local resources effectively and guard against natural hazards, we must

observe and understand the Earth as a whole system.”116

Much of the PEI statement’s language resembles that used by the ESSC. The statement,

submitted two months after the publication of the ESSC’s Closer View report in January 1988,

explicitly linked PEI to other research efforts, including NASA’s “Earth System Sciences

studies” and recommended cooperation amongst these endeavours. However, the PEI statement

made it clear that, to understand global problems—like ozone depletion, carbon dioxide buildup

in the atmosphere, ocean pollution, and deforestation—it was necessary to collect, “observations

of the whole Earth, and to view such changes on a time scale of decades, centuries, or even

geologic time.” Timescales beyond decades to centuries—geological timescales of thousands to

millions of years—were also important. According to the PEI statement, “The earth’s natural processes involve not only the land, oceans, ice-covered polar regions and atmosphere, but also the deep interior of the earth and the solar-terrestrial environment.”117 In 1989, AGU President

Don Anderson similarly described PEI as transcending the ESS research program: “no part of the

Earth can be understood in isolation from other parts and...‘global change’ is not just a problem of the fluid envelopes of the Earth. In short, we are proposing an integrated study of our planet,

115 American Geophysical Union, Planet Earth Initiative, Position Statement, [1988], Folder 11, Box 20, AGU Records, AIP, College Park, MD. 116 Planet Earth Initiative: A Recommendation from the American Geophysical Union, [1988], Folder 11, Box 20, AGU Records, AIP, College Park, MD [emphasis added]. 117 Ibid.

318

from a basic research point of view.”118 What the AGU proposed was to develop a fundamental

understanding of the Earth, which could only be achieved by incorporating processes on all

timescales into the research. With its primary emphasis on developing a computational model of

the Earth system, the ESSC had to limit its focus for practical purposes and, therefore, chose to focus on timescales of decades to centuries. This made the ESSC’s ESS incomplete and, therefore, less appealing for many geophysicists. The AGU’s PEI proposal could claim to study the “whole Earth” because what it took to be a fundamental understanding of the Earth did not involve building a computational model of it.

The AGU sought to promote a research program that went beyond the ESSC’s Earth system science while still supporting a generic Earth system concept that entailed the Earth being

understood and studied as a system. Unfortunately, like the Earth-as-a-System Committee, the

PEI did not achieve the desired effects. Early criticism of the draft PEI Statement suggested that,

“The position statement is broad and not specific….It describes a general research program as opposed to a program that creates the ability to do research.”119 Regarding the finalized version

of the PEI statement, chairman Cicerone wrote to his fellow panel members expressing

satisfaction with their product, even if, “Our critics could still ask for a stronger, more specific

punch line, i.e. detail in the recommendation.” Cicerone believed that their work was meant to

set the stage for more global cooperation, and recommendation details would be filled in later by

scientists and politicians.120 The PEI had accomplished more than the “almost nothing” of the

Earth-as-a-System Committee, but not much more.

118 Letter from Don L. Anderson to P. Vanicek, 15 Aug 1989, Folder 11, Box 20, AGU Records, AIP, College Park, MD [emphasis added]. 119 Planet Earth Initiative (PEI) Draft Report, [1988], Folder 13, Box 29, AGU Records, AIP, College Park, MD. 120 Memorandum from Ralph J. Cicerone to Dave Cauffman, et al., 24 Feb 1988, Folder 13, Box 29, AGU Records, AIP, College Park, MD.

319

In an effort to address the multiple and manifest deficiencies of the Earth-as-a-System

Committee (and the parallel PEI), AGU officials reconstituted the Earth-as-a-System Committee into the Planet Earth Committee (PEC) in May 1988.121 PEC would take over and expand on the

duties of the previous committee by further developing and promoting PEI. AGU members

reported enthusiastic responses from politicians and government officials for PEI. The challenge

now was to build on this enthusiasm by developing a concrete research strategy.122 PEC would

do this by attempting to, “further define the Planet Earth Initiative.”123 AGU official Les

Meredith later wrote of the “deficiencies” of PEI, including its failure to outline a “program” that

could be used to “sell” it. Meredith colloquially suggested that, “PEC was supposed to add this

one level of meat to the PEI bones.”124 AGU Council meeting minutes from December 1989

stated that PEC should “amplify” the PEI statement.125 According to Don Anderson, PEC

should go beyond PEI’s focus on “climatic change, disasters, and resources” to include all basic

research on the Earth “as a planet” with the ultimate goal of developing a general

“understanding” of the Earth.126 PEC would “define and help get implemented” PEI with

activities that included: working with AGU members and other scientific societies to develop

PEI; organizing special sessions, meetings, and publications to assemble the views of the

research community for PEI’s content; educating AGU members, decision makers, and the

121 Action Items, [AGU] Executive Committee Meeting, 15 May 1988, Folder 10, Box 14, AGU Records, AIP, College Park, MD. 122 [AGU] Executive Committee Minutes, 15 May 1988, Folder 7, Box 13, AGU Records, AIP, College Park, MD. 123 Action Items, [AGU] Executive Committee Meeting, 15 May 1988, Folder 10, Box 14, AGU Records, AIP, College Park, MD. 124 Letter from L.H. Meredith to William Kaula, 29 Nov 1989, Folder 9, Box 20, AGU Records, AIP, College Park, MD. 125 Minutes of [AGU] Council Meeting, 5 Dec 1989, Folder 1, Box 11, AGU Records, AIP, College Park, MD. 126 Telemail from D. Anderson to F. Spilhaus, 28 Feb 1989, Folder 12, Box 20, AGU Records, AIP, College Park, MD.

320

general public on PEI; and, continually updating PEI recommendations as “progress is

achieved.”127

PEC’s first major effort was a two-page writeup that explained in “layman’s language” how PEI should be implemented. Les Meredith shared a draft of this with PEC members in

November 1989. With this document, PEI now had a more focused objective: to take measurements of the Earth’s “basic properties” using a global observing system and associated . Linking this work to other programs that studied the Earth (e.g. the IGBP, WCRP, and the Global Change Research Program), the idea was that PEI research would provide a common baseline of measurements that could be used by other programs, thus increasing their

comparability, efficiency, and comprehensiveness. This “unified approach” was, according to

the PEC two-pager, being used to study other planets, “and would undoubtedly have already

been done with the Earth if we had started with existing technology and a clean slate instead of

with the somewhat random way in which studies of Earth have evolved over the whole of human

history.” PEC did not call for the measurement of all of the Earth’s “almost limitless”

components. This was “neither necessary nor desirable.” Instead, PEC recommended focusing

on a core group of essential measurements: key forces (solar radiation, Earth’s interior structure,

human environmental modifications) and the primary properties that comprise the Earth’s

atmosphere, oceans, land, and biology (cloud cover and type, atmospheric composition and

motion, ocean temperature and motion, ice depth and extent, land water availability, land surface

composition and motion, surface heat fluxes).128

127 Planet Earth Committee write-up, [1989], Folder 12, Box 20, AGU Records, AIP, College Park, MD. 128 [Draft Two-Pager] Planet Earth Program of the Planet Earth Initiative, 22 Nov 1989, Folder 9, Box 20, AGU Records, AIP, College Park, MD.

321

The PEC two-pager hypothesized that these essential measurements would be fundamental for all current and future research programs focused on studying the Earth’s “global problems” (e.g. ozone depletion, climate change, droughts, earthquakes). These measurements would be needed to, “first determine the present state of that portion of the total Earth system

with which they’re concerned.”129 Meredith underscored this point when he forwarded the draft

two-pager to PEC members: “In summary, the American Geophysical Union firmly believes that

looking at the Earth as one total system is important both scientifically and to our future well

being.”130 Just as with its earlier work, this AGU group drew on the “Earth system” language

popularized by the ESSC for describing an interconnected planet. However, PEC defined the

Earth system altogether differently than the ESSC. PEC’s Earth system was not limited to the

shorter term processes of the fluid Earth but incorporated those longer term processes occurring

in the solid Earth. PEC adopted the ESSC’s Earth system concept without fully endorsing the

ESSC’s specific Earth system definition or its ESS research program based on that definition.

A “Complete” Mission to Planet Earth

Like the Earth-as-a-System Committee and PEI, PEC achieved little in the long term. (That the

ESSC achieved more success might be a testament to the management skills of Francis

Bretherton, Laura Lee McCauley, and other support staff, as well as to the committee’s

consequential use of email rather than regular mail to communicate.) Despite this lack of

success, an examination of PEC’s major activity—the preparation of a report intended to provide

policy-makers and AGU members more scientific details about PEI—reveals the depth of the

129 Ibid. 130 Telefax cover sheet from Les Meredith to Planet Earth Committee, 22 Nov 1989, Folder 9, Box 20, AGU Records, AIP, College Park, MD [emphasis added].

322

AGU’s commitment to viewing and studying the Earth as a system. It demonstrates how readily

AGU members adopted the Earth system concept, even while they tried to transform the ESSC’s

Earth system science research program into something more comprehensive, something that centered not around constructing a computational model of the Earth system. Instead, PEC developed a broad research program that focused on assembling “fundamental knowledge” about the entire planet and its environs. With this shift in focus, PEC did not need to limit its focus to a specific timescale for practical purposes, but could incorporate research focused on all timescales, short or long. PEC called its proposed research program a “Whole ‘Mission to Planet

Earth’” and later a “Complete ‘Mission to Planet Earth.’”131 It would be “complete” in a way

that, according to AGU members, the ESSC’s ESS research program could never be, with its

narrow focus on shorter term processes that marginalized geophysical research. This report

never got beyond the draft stage, but its preparation exposes the concerns of geophysicists as

they both supported the ESSC’s ideas regarding the Earth as a system but also fought to ensure

that ESS was surpassed by something more comprehensive.

PEC’s “Complete ‘Mission to Planet Earth’” emerged out of a PEC-organized AGU

convention session held in the spring of 1989.132 Titled “Earth: A Changing Planet,” the purpose

of the session was to detail major global problems in the Earth sciences, examine the

technological prospects of carrying out research in these areas, connect this work to needs in

resource, hazard, and environmental management, and relate these scientific and social needs to

ongoing or proposed research programs. Four invited speakers provided the core scientific

material for the session: geologist Peter Wyllie spoke on “The Earth’s Interior and Crust”;

131 Telefax cover sheet from Les Meredith to Planet Earth Committee, 22 Nov 1989, Folder 9, Box 20, AGU Records, AIP, College Park, MD; A Complete “Mission to Planet Earth”: A Policy Statement of the American Geophysical Union, December 1989, 4 Nov 1989, Folder 10, Box 20, AGU Records, AIP, College Park, MD. 132 Agenda Item for AGU Council Meeting December 5, 1989: Planet Earth Committee, by Adam Dziewonski, chairman, Nov 1989, Folder 9, Box 20, AGU Records, AIP, College Park, MD.

323

atmospheric scientist and ESSC member Ronald Prinn on “Atmosphere, Oceans, and Land

Surface: Our Physical Environment”; biologist and meteorologist (and future NASA astronaut)

Piers Sellers on “Biosphere Interactions”; and, geologist M. Gordon Wolman on “The Impact of

Man.” Each speaker provided a brief overview of the state of research in their area, the new

techniques or technologies that offered fruitful avenues of future research, and the observational

programs that would best serve this research. As the titles indicate, research topics ranged from

shorter term processes in the atmosphere, oceans, and on land to the longer term processes

occurring in the Earth’s core and mantle.133

These speakers drew on the ESSC’s systems language to describe the Earth as an object

of scientific research, but were careful not to align themselves with the ESSC’s Earth system

science research program. Speakers at the “Earth: A Changing Planet” session did not want to

place research limits based on timescales as the ESSC did. Peter Wyllie’s presentation forcefully

illustrated this. It focused specifically on longer term Earth processes occurring in the Earth’s

core and mantle, and described research currently underway to study solid Earth science topics

like the geochemistry, physical state, and dynamics of the mantle and core, and the subduction

and continental collisions of lithospheric plates. Wyllie noted the emergence of biogeochemical

cycles as a “major theme in recent years,” but he worried that this research placed too much

emphasis on the cycling that occurred in the Earth’s outer layers, in the atmosphere, oceans, and

on land surfaces. Wyllie implored listeners, “not [to] lose sight of the deeper part of these

cycles, such as the important role of subduction and volcanism in circulating and remobilizing

H2O, C, N, P, S, etc.” Rather than overemphasize satellite observations as some claimed the

ESSC had done, Wyllie described a more balanced research program that emphasized the role of

133 Memorandum from A.M. Dziewonski to Members of Planet Earth Committee, et al., 17 Aug 1989, Folder 11, Box 20, AGU Records, AIP, College Park, MD.

324

both space-based and in-situ measurements in a single, integrated observational system.

NASA’s EOS would include instruments to map the Earth’s magnetic and gravity fields, and the

movement of the lithosphere’s plates. A Permanent Large Array of Terrestrial Observatories

(PLATO) would include in-situ measurements taken from ocean floor stations, the global

seismic network of sensors, GPS receivers, and drilling operations. In case the subtle critique of

the ESSC was lost on any listeners, Wyllie bluntly stated that, “It is clearly time to develop

‘Earth System Science’ for the whole earth[.]” It was, according to Wyllie, time to transcend the

timescale limits and emphasis on space observations imposed by the ESSC. It was time to

develop a more comprehensive “Earth System Science” that included the whole Earth, especially

those parts of the Earth studied by AGU members. Wyllie offered this critique of ESS but still

unreflectively expropriated the ESSC’s Earth system concept. He emphasized geological and

geophysical processes, but went on to notice that global issues like the greenhouse effect and

ozone depletion in the “earth system are much on our minds these days[.]”134 Session presenters

uncritically adopted the Earth system concept while at the same time rejecting what they

perceived to be the ESSC’s overly narrow ESS research program.

This “very successful” session motivated PEC to turn the presentations into a “brochure”

that would more systematically and consistently describe an Earth science research program for

the “whole” planet. PEC chair and seismologist Adam Dziewonski reported that, at a subsequent

PEC meeting, it was decided that the best way to expand and promote PEI was to prepare a

“brief brochure” of around 30 pages to describe the proposed program, aimed primarily at

federal agencies and other federal decision makers.135 According to Les Meredith, “For the

134 [Peter] Wyllie, “The Earth’s Interior and Crust,” [May 1989], Folder 11, Box 20, AGU Records, AIP, College Park, MD [emphasis added]. 135 Memorandum from Adam Dziewonski to Planet Earth Committee, et al., 17 Aug 1989, Folder 11, Box 20, AGU Records, AIP, College Park, MD.

325

document to be useful, it must be easily understandable by people on the Hill. That is, by

average college graduates” without Earth science training.136 The four primary scientific

speakers at the session (Wyllie, Prinn, Sellers, and Wolman) were asked to prepare 10-page

summaries of their presentations and to point out what would be the most exciting research over

the next 20 years. These chapters (called: solid earth, oceans and atmospheres, biosphere interaction, and anthropogenic effects) would, along with a writeup on the “Global Observing

System,” convey PEI ideas to the intended audiences. According to Dziewonski, “The purpose of the brochure would be to establish the point that the necessary step in achieving the understanding of the workings of Planet Earth is the setup of a multidisciplinary global observing system with observations carried out both from space and on the Earth’s surface.”137 This

ostensibly would be a study of the whole planet, and not just the ESSC’s Earth system with its

putatively narrow timescale boundaries. As Wyllie noted, geophysicists like himself believed

that there was a need to, “develop Earth System Science for the whole earth.” ESS might study

important components of the Earth system, but Wyllie maintained that it was missing a

component in that interactive system, the “geosphere.”138 Despite this perceived lacunae, Wyllie

(and his fellow presenters, PEC members, and the AGU more broadly) still relied on the ESSC’s

Earth system concept to describe the Earth as an object of scientific study.

For its “brochure,” PEC drew heavily from the presentation session as well as from two

Earth science reports prepared by groups within the NAS. First, there was the Committee on

Earth Sciences (CES) report Our Changing Planet: A U.S. Strategy for Global Change Research

136 Letter from Les Meredith to William Kaula, 30 Aug 1989, Folder 11, Box 20, AGU Records, AIP, College Park, MD. 137 Memorandum from Adam Dziewonski to Planet Earth Committee, et al., 17 Aug 1989, Folder 11, Box 20, AGU Records, AIP, College Park, MD. 138 [Peter] Wyllie, “The Earth’s Interior and Crust,” [May 1989], Folder 11, Box 20, AGU Records, AIP, College Park, MD [emphasis added].

326

(1988) which outlined an, “intense interagency effort by experts in various earth sciences and

other disciplines” to develop goals, an implementation strategy, and a research budget for the US

Global Change Research Program (USGCRP).139 The USGCRP’s scientific objectives outlined

in the report would be to monitor, understand, and eventually predict global changes,

“encompassing the full range of earth system changes, including climatic, seismic, ecological,

and biological changes.”140 It divided Earth science research into seven interdisciplinary areas:

(1) Biogeochemical Dynamics, (2) Ecological Systems and Dynamics, (3) Climate and

Hydrological System, (4) Human Interactions, (5) Earth System History, (6) Solid Earth

Processes, and (7) Solar Influences. Our Changing Planet broke with ESS’s focus on shorter

timescales, but it still adopted the ESSC’s Earth system language to refer to the planet as an

interconnected scientific object. The phrase “Earth system” appeared 35 times in the report

body’s 40 pages. Given CES’s proposal for a more expansive Earth science program and its

direct links to the federal government, the AGU interpreted Our Changing Planet as being the

federal government’s “plan” for Earth sciences, and internal correspondence suggested that it

should be used as “input for the PEC fleshing out activities.”141

The second source for PEC’s “brochure” was the SSB’s Task Group on Earth Sciences

report Mission to Planet Earth: Space Science in the Twenty-First Century (1988).142 Chaired

by the AGU’s Don Anderson, membership on this Task Group was comprised almost entirely of

either former ESSC members (James Baker, Moustafa Chahine, Berrien Moore, Ronald Prinn) or

139 Erik Conway states that the Bush Administration launched the USGCRP in 1989 to better determine the planetary effects of increasing levels of carbon dioxide in the atmosphere. See: Conway, Atmospheric Science at NASA, 6. 140 Committee on Earth Sciences, National Research Council, Our Changing Planet: A U.S. Strategy for Global Change Research (Washington, DC: National Academies Press, 1988), 3 [emphasis added]. 141 [Les Meredith] to Don [Anderson], [Feb 1989], Folder 12, Box 20, AGU Records, AIP, College Park, MD. 142 Task Group on Earth Sciences, Space Science Board, National Research Council, Mission to Planet Earth: Space Science in the Twenty-First Century: Imperatives for the Decades 1995 to 2015 (Washington, DC: National Academy Press, 1988).

327 current PEC members (Dziewonski, William Kaula, Donald Turcotte). That PEC drew on this document for its brochure is, therefore, not surprising. The SSB formed this Task Group, at the behest of NASA, to determine the main Earth science issues that might be addressed using space technologies like Earth observing satellites. Again using the “Mission to Planet Earth” phrase and describing the interconnected planet as the “Earth system,” the report outlined a unified research program that called for studying the Earth “from its deep interior to its fluid envelopes” using both space-based and in-situ observations. The project revolved around what its authors called four “grand themes”:

1. To determine the composition, structure, dynamics, and evolution of the Earth's

crust and deeper interior.

2. To establish and understand the structure, dynamics, and chemistry of the

oceans, atmosphere, and cryosphere, and their interaction with the solid Earth.

3. To characterize the history and dynamics of living organisms and their

interaction with the environment.

4. To monitor and understand the interaction of human activities with the natural

environment.

To research these four grand themes, the authors recommended data collection with a global observation system with five large geostationary satellites, two to six large polar-orbiting platforms, a series of smaller satellites flying specialized missions, and PLATO observatories that would collect in-situ measurements at hundreds, potentially thousands, of sites. This far- ranging research was exactly what PEC members wanted to endorse. PEC-related correspondence at the AGU repeatedly emphasized the need to draw on the four grand themes of

328

the Mission to Planet Earth report.143 That all four of the presentations at PEC’s “Earth: A

Changing Planet” session closely aligned with Mission to Planet Earth’s four themes was no

coincidence. They were expressly designed to do so.144

The SSB’s Mission to Planet Earth report mentioned the findings of other committees

and their research programs, including the ESSC’s ESS, but stressed that its proposal superseded

all previous efforts. Chairman Don Anderson explicitly noted that the Mission to Planet Earth

report, “has gone far beyond them [previous committees and reports] to develop a broader, more

comprehensive, and long-range study of Earth for the twenty-first century.”145 In March 1989,

in his capacity as AGU President, Don Anderson testified before the Senate Committee on

Commerce, Science and Transportation on this Mission to Planet Earth. Anderson emphasized a

lacuna in Earth science knowledge, observing that: “There has never been a systematic study of

our own planet. We have mounted ambitious expeditions to other worlds and, in many respects,

we now understand them better than we understand the Earth.” According to Anderson, this

Mission to Planet Earth program was the first to recommend an integrated study of the entire

planet: “in order to understand the Earth you have to study the whole Earth.”146

The “whole Earth” phrase was an echo of PEI’s earlier proposal and perhaps also the

AGU’s ambition to study the Earth as a “total system.” Either way, Anderson’s testimony was a

rejection of ESS’s shorter timescales which Anderson perceived as being at odds with AGU

interests. From the vantage point of the geophysics community, ESS was not comprehensive

143 Notes for Planet Earth Committee File, 11 May 1989, Folder 12, Box 20, AGU Records, AIP, College Park, MD; Mission to Planet Earth, testimony of Don L. Anderson to the Committee on Commerce, Science, and Transportation of the United States Senate, 8 Mar 1989, Folder 11, Box 20, AGU Records, AIP, College Park, MD. 144 [Peter] Wyllie, “The Earth’s Interior and Crust,” [May 1989], Folder 11, Box 20, AGU Records, AIP, College Park, MD. 145 Task Group on Earth Sciences, Mission to Planet Earth, xi. 146 Mission to Planet Earth, testimony of Don L. Anderson to the Committee on Commerce, Science, and Transportation of the United States Senate, 8 Mar 1989, Folder 11, Box 20, AGU Records, AIP, College Park, MD [emphasis added].

329

enough. More was required. However, like Our Changing Planet before it, Mission to Planet

Earth frequently and unreflectively used the ESSC’s Earth system phrase to refer to the planet as

a scientific object. It even included an entire chapter that discussed the Earth “as a system.”147

Clearly, this concept was not, for the report’s authors, tied to the ESSC’s specific ESS research program, but simply gestured at an interconnected planet that required interdisciplinary study. In this broad, ill-structured sense, there was no dissent. The Earth system concept was being embraced by a wide variety of Earth scientists. It was only when specialists sought to more rigidly define the contours of the Earth system that disagreements arose. The multivalent character of the Earth system made it useful.

The Mission to Planet Earth report recommended a research program largely in line with the comprehensive study of the Earth envisioned by the AGU, but no one was actually promoting the adoption of the program. PEC could, the AGU believed, provide this service. An early draft of the PEC’s mandate noted that the Mission to Planet Earth report outlined a potential national research program on global issues, but that, “no organization or agency is following up to either sell the program to public officials and agencies or to flesh out the program with detailed sciences.” The strategy was that PEC could “fill this void and make the program its own.” PEC was the group that should attempt to place a “Mission to Planet Earth” on the “National agenda along with the SSC [Superconducting Supercollider], Human Genome [Project] and [Hubble]

Space Telescope.” Major projects like these “didn’t just happen” but had to be vigorously developed and then “sold” to scientists, politicians, and the general public.148

147 Task Group on Earth Sciences, Mission to Planet Earth. 148 Telemail from D. Anderson to A. Dziewonski and B. Dalrymple, 1 Feb 1989, Folder 12, Box 20, AGU Records, AIP, College Park, MD.

330

From the beginning of PEC’s “brochure”—which swelled from the expected 30 pages to over 80 pages—it was clear that PEC’s research program was billed as being more encompassing than that offered by the ESSC. Called “A Complete ‘Mission to Planet Earth,’” the authors used the “Earth system” as if it were something already accepted as a fact about the Earth that required no justification. The Earth was, for the authors, already a system. They did this despite their lack of enthusiasm for the ESSC’s Earth system science research program that was perceived as too narrowly focused on shorter timescales that marginalized geophysical research.

The authors noted that the Earth had, “not been exposed to systematic study for a variety of logistical, economic, and political reasons.” They further claimed that there was “growing awareness” of the need to study the planet as an “integral whole” that would include shorter term problems in the “fluid parts” of the planet (the atmosphere and oceans) along with longer term problems like those in the Earth’s interior. The last draft149 contained science chapters on the

“Interior and Crust” by Wyllie, “The Sun and the Earth” by A.F. Cheng, “Atmosphere, Oceans,

and Land” by Prinn, “Biosphere Interactions” by Sellers, and “Impact of Man” by Wolman.

These chapters represented only a slight expansion on the four “grand themes” of the SSB’s

Mission to Planet Earth report and PEC’s “Earth: A Changing Planet” presentation session. The

addition was to examine the relationship between the sun and the Earth. To improve knowledge

of these (new) five “grand themes,” the authors made six recommendations: (1) Develop a global

observation system of the Earth’s environmental properties; (2) Develop this observation system

using existing technologies and without subservience to other programs or needs; (3) Develop a

commensurate data management system and modeling capabilities to utilize and analyze

observational data; (4) Make the effort international; (5) Immediately take certain actions

149 There is no evidence that the PEC’s “Complete ‘Mission to Planet Earth’” ever made it beyond the draft stage.

331

relating to environmental quality (e.g. reduce carbon dioxide and CFC emissions, develop fuel-

efficient vehicles and alternative energy sources, halt forest destruction, and minimize/safely

dispose of waste); and, (6) Educate the American public about environmental and resource

issues.150 Whatever the specific merits of PEC’s brochure, there could no longer be any doubt

that this was intended to be a more expansive research project than what was offered by the

ESSC.

This “Complete ‘Mission to Planet Earth’” was perceived as complementing “other

studies” and program proposals, including ESS. However, this complete mission should,

according to AGU official Les Meredith, “fill their voids and weaknesses.”151 As

Recommendation Two of the report stated, this “Complete Mission to Planet Earth” should not

be subjected to the whims or needs of other programs, presumably referring to ESS developed by

a NASA-affiliated group that prioritized shorter term processes, space-based research, and

computer modeling. The report’s introduction cautioned that, though the phrase “mission to

planet earth” had originated in work that led to the Mission to Planet Earth report, NASA had

subsequently “adopted” the phrase and now it connoted a “space-based approach” to studying the planet: “Hence this document adds ‘Complete’ to its title to emphasize a return to a more comprehensive and balanced approach.” It was then hinted that other similar reports were biased in a way this one was not. Other reports had been, “funded by other bodies, whose interests may have influenced, subconsciously if not consciously, the recommendations.”152 This was an odd

claim given that Ronald Prinn had contributed to this introduction and was also a member of the

150 A Complete “Mission to Planet Earth”: A Policy Statement of the American Geophysical Union, December 1989, 4 Nov 1989, Folder 10, Box 20, AGU Records, AIP, College Park, MD. 151 Letter from Les Meredith to William Kaula, 30 Aug 1989, Folder 11, Box 20, AGU Records, AIP, College Park, MD. 152 A Complete “Mission to Planet Earth”: A Policy Statement of the American Geophysical Union, December 1989, 4 Nov 1989, Folder 10, Box 20, AGU Records, AIP, College Park, MD.

332

ESSC. Since the AGU worked only to promote geophysical research and not specific research

methods (i.e. satellite data collection and computational modeling), the implication here was that

it could be focused wholly on best scientific practices rather than on justifying expensive space

technologies like the Space Shuttle or large satellite platforms like NASA’s EOS. This

“complete” research would not be focused on modeling the fluid atmosphere and oceans where

Earth processes occurred on shorter timescales. As the chapter on the “Interior and Crust” stated

(taking a line directly from Wyllie’s presentation), “It is clearly time to develop ‘Earth System

Science’ for the earth’s crust and interior, following the example of those mainly concerned with atmospheres and oceans.” It was, for PEC, clearly time to develop a comprehensive Earth system science that protected AGU interests in a way that ESS had not.

While rejecting the ESSC’s specific project proposal, the “Complete ‘Mission to Planet

Earth’” appropriated the language and ideas of the ESSC, referring to the planet as the “Earth system” and noting the need to study the Earth “as a system.”153 These references are frequent

and unreflective. There was no question about whether or not there was such a thing as the

“Earth system.” (There was also no sensitivity to the way the ESSC acknowledged that the

“Earth system” concept had been specifically constructed to meet the modeling needs of ESS.)

For the authors of the “Complete” report, the Earth system was implicitly accepted as a boundary

object between different Earth science communities. It was adopted by a variety of committee

members that largely agreed on its expansive, ill-structured meaning—that the Earth was an interconnected system requiring interdisciplinary study—even if they disagreed on how to specifically define the Earth system. PEC and the ESSC could both use the “Earth system” concept without meaning precisely the same thing. PEC and the AGU clearly did not favour

153 Ibid.

333

defining the Earth system by timescales of decades to centuries as the ESSC had done and they

did not think that the “challenge” of studying the “Earth system” required modeling it in its

entirety. This might have relegated much geophysical research to the margins for practical

reasons. But the Earth system concept demanded no such specific agreement about its meaning.

It could be adopted by very different practitioners, even as it was given more specific meanings

by specialists in local contexts. In such contexts, the main question was not whether or not the

Earth system existed, but what components were thought to comprise this system.

Hence the ongoing overemphasis of the word “complete” at the beginning of PEC’s

brochure (and words like “whole” and “total” in other AGU-motivated reports). This was to be a complete mission to planet Earth. It was so critical for PEC to emphasize that their report would go beyond the work of the ESSC that geodicist William Kaula (the brochure’s primary author and editor) brainstormed other options to emphasize that the research being promoted by PEC would incorporate all of the Earth. A gamut of possible synonyms was tabled: entire, whole, total, comprehensive, uncut, unabridged, thorough, fundamental, real, genuine, essential, perfect, ideal, and exemplary.154 Whatever the exact word chosen, the intention was clear. The AGU’s

program outlined by PEC would be a more comprehensive study of the Earth than the program

proposed by the ESSC if only in its tendency towards expansive disciplinary inclusivity. In

October 1989, PEC prepared a draft letter to Allan Bromley—Assistant to the President for

Science and Technology—on the proposed Global Change Research Program. In line with the

desire for a “complete” study of the planet, the letter once again expressed PEC’s concern with,

“the lack of a coherent scientific program addressed to understanding the Earth in all its

regimes—solid, liquid, gaseous, plasma and biota—over the entire range of spatial and temporal

154 Notes on “A Complete ‘Mission to Planet Earth,’ September 1989, Draft No. 1” by Don Anderson, 9 Sep1989, Folder 11, Box 20, AGU Records, AIP, College Park, MD.

334

scales.” It listed a series of recommendations that would enable the US to “take the lead” in the

collection and management of observations on all of the Earth’s components. And once again

the letter emphasized that, “the American Geophysical Union firmly believes that looking at

Earth systematically is important to science and...to the future well-being of mankind.”155

Though PEC’s stated aim was to provide a more complete research program than

previous efforts, some reviewers took issue with early drafts of the brochure for not in fact being

“complete.” Don Anderson emphasized that early brochure drafts did not incorporate the upper

atmosphere and near-space environs and, therefore, did not fulfill the call for a “complete”

mission to study the Earth. He identified two areas—the ionosphere and magnetosphere—that were largely absent from the August 1989 brochure draft. According to Anderson, this “Planet

Earth” that was to be studied here was, “not [a] complete Earth yet[.]”156 AGU member and space physicist Christopher T. Russell was forwarded a copy of the draft brochure by a

“concerned” AGU member. As Russell reported, “His concern is, and I concur, that the

‘Complete mission’ is not complete. To restrict one’s concerns to the narrow boundary layer in which man maintains his physical presence is a rather narrow viewpoint.” Russell recommended the addition of a chapter that covers the upper atmosphere, ionosphere and magnetosphere.157

These concerns prompted the addition of Cheng’s “The Sun and the Earth” chapter to the final draft, which incorporated aspects of space and plasma physics into the overall research program.158 That PEC was willing to add a completely new chapter at a late date to address

155 Draft of Letter to Allan Bromley from Don Anderson, prepared by the Planet Earth Committee, [1989], Folder 1, Box 11, AGU Records, AIP, College Park, MD [emphasis added]. 156 Notes on “A Complete ‘Mission to Planet Earth,’ September 1989, Draft No. 1” by Don Anderson, 9 Sep1989, Folder 11, Box 20, AGU Records, AIP, College Park, MD. 157 Letter from C.T. Russell to D.L. Anderson, 19 Sep 1989, Folder 11, Box 20, AGU Records, AIP, College Park, MD. 158 A Complete “Mission to Planet Earth”: A Policy Statement of the American Geophysical Union, December 1989, 4 Nov 1989, Folder 10, Box 20, AGU Records, AIP, College Park, MD.

335

these concerns suggests that it was genuine about developing as complete a program as possible.

Disciplinary inclusivity was a high priority.

In November 1989, Les Meredith raised another issue with PEC regarding the brochure’s

“completeness.” Meredith claimed that it was still, “not a complete ‘Mission to Planet Earth.’ In

fact, it is just the reverse. Both the Grand Themes and the Recommendations are much more

narrowly focused than either PEI or the NAS ‘Mission to Planet Earth.’” Meredith provided few

details, but his worry seems to have been that the recommendation to “Develop global systems

for environment and resources only” was “contrary to looking at the Earth as a total system.”159

According to Meredith, an observation system that focused too narrowly on only resource and

environmental issues would not provide the more comprehensive planetary research program

desired by Meredith and other AGU members. Kaula responded by complaining that he was

unclear what “significant things” had been left out of the “‘Grand Themes and

Recommendations’ that make them more narrowly focussed.” Kaula, perhaps now experiencing

some of Bretherton’s pains while chairing the ESSC, explained that to make the program

compelling it was necessary to emphasize the most important aspects. The categories of

“environment and resources” cover many aspects of the Earth, and therefore should not properly

be considered “narrow.”160 Meredith countered: “We seem to have a basic disagreement of what’s ‘most important.’ PEI says it’s to understand the Earth as one total system. You seem to feel that there must be priorities within the program and that these must stress societal as opposed to intellectual activities….I don’t think it is the thrust of either the PEI or of ‘Mission to

159 Memorandum from Les Meredith to Planet Earth Committee and Writing Subcommittee, 17 Nov 1989, Folder 10, Box 20, AGU Records, AIP, College Park, MD [emphasis added]. 160 Letter from William M. Kaula to L.H. Meredith, 27 Nov 1989, Folder 9, Box 20, AGU Records, AIP, College Park, MD.

336

Planet Earth.’”161 This disagreement is reminiscent of the early days of the ESSC. Meredith wanted to ensure that the brochure promoted a fundamental or “intellectual” study of the planet, rather than a research program that emphasized social or policy utility and, therefore, made scientific prioritizations. All basic Earth science research on the planet was equally important.

Making priorities risked doing something like what the ESSC had done. For practical purposes, the ESSC chose to emphasize processes occurring on timescales of decades to centuries in part because that scale of change was most relevant to humans. PEC’s brochure was intended specifically to avoid this kind of scientific ranking that had the potential to marginalize geophysical research areas.

What is important here is not the actual effectiveness of the Planet Earth Committee and its brochure (or the Earth-as-a-System Committee or the Planet Earth Initiative). All of these

AGU activities might appear convoluted and even peripheral. In a certain sense they were, since they had many overlapping aims, unclear objectives, and none explicitly contributed to government policy about Earth science research or the study of global change. PEC was formed to address the deficiencies of the Earth-as-a-System Committee and PEI. But PEC’s “Complete

Mission to Planet Earth” brochure never got beyond the draft stage. The AGU’s Council discussed the final draft brochure at its fall 1989 meeting. Members felt that parts of the document could be used in AGU policy documents and testimony, but ultimately concluded that the document was too general and that it tended to venture beyond the AGU’s focus on geophysical issues into the territory of “economic, social, and political issues.” It was, therefore,

161 Letter from L.H. Meredith to William Kaula, 29 Nov 1989, Folder 9, Box 20, AGU Records, AIP, College Park, MD.

337

“not approved” by the Council.162 PEC, like the Earth-as-a-System Committee and PEI before it, achieved little of lasting substance.

Regardless of the effectiveness of these AGU activities, it is noteworthy that AGU members were having these kinds of conversations about forming a specific project to study the

Earth system. The existence of these committees and initiatives demonstrate that AGU members were concerned with promoting more systematic studies of the planet as suggested by the ESSC.

They adopted the ESSC’s vocabulary and ideas, even while maintaining that the ESSC’s specific

ESS research program was too limited. The AGU readily adopted the Earth system concept as a way to refer to an interconnected planet requiring interdisciplinary Earth and environmental science research. They did this while rejecting ESS because its focus on timescales of decades to centuries was perceived as neglecting the interests of the geophysical research community

(particularly that research focused on the Earth’s core, mantle, and outer environs). The AGU’s repeated claim was that it was necessary to fully understand the whole, total, or complete Earth as a system. Members also suggested that ESS, as a product of a NASA-affiliated committee, was overly focused on space-based research to the detriment of in-situ studies. Even though it was never formally adopted, the draft brochure and the work of PEC show that the AGU readily embraced the Earth system concept promoted by the ESSC while rejecting the ESSC’s specific definition of the Earth system. The AGU’s Earth system needed to be more “complete” than the

ESSC’s. AGU members argued that the total Earth system included longer term processes and extended from the core through the magnetosphere to the sun. These needed to be studied alongside shorter term processes in order to have a comprehensive understanding of the Earth.

That the AGU could adopt the Earth system concept without agreeing with how the ESSC

162 Letter from William M. Kaula to A.F. Spilhaus, 8 Nov 1989, Folder 9, Box 20, AGU Records, AIP, College Park, MD; Minutes of [AGU] Council Meeting, 5 Dec 1989, Folder 1, Box 11, AGU Records, AIP, College Park, MD.

338 defined the Earth system confirms that the concept has utility, not despite of, but because of, its vagueness. The plurality of meanings meant that specialists did not have to agree with other specialists about the Earth system’s precise definition in disciplinary settings. The Earth system was a boundary object that functioned in specific ways in specialized contexts, but it also had a broader meaning that facilitated communication across disciplines and institutions. It could be used by the ESSC, the AGU, and other Earth scientists without requiring agreement on specific details.

CONCLUSION

Perhaps counterintuitively, vagueness can sometimes be a strength. The word “vague” need not be pejorative. To be vague is to be flexible, capable of multiple interpretations, of being used differently by different individuals and groups, but with enough commonality that cooperation or collaboration or at least communication remains possible. Of course, vagueness can be a problem. Much depends on where exactly the vagueness lies. “Gaia,” for instance, could not become a boundary object like the “Earth system” because it is problematically vague. One meaning of Gaia for some Earth scientists is almost trivially true (e.g. Earth is a planet where organisms contribute to planetary conditions) while for others, another meaning of Gaia is highly dubious and unscientific (e.g. Earth is a planet where organisms intentionally optimize their environment for survival). If a scientific concept can avoid the controversy, it can have great utility and, therefore, mobility. The term “planet” is a case in point. Originally, different groups defined the word “planet” in a variety of ways that differed in detail but contained enough similarities to facilitate broad communication. After the IAU voted to formalize its definition,

“planet” lost its vagueness and, as a result, its ability to facilitate inter-group communication.

339

In the sciences, when vagueness is useful, the vague concept or word or image or idea or

object can become a “boundary object” that links together different communities. The Earth

system is such a boundary object. It was the phrase chosen by the ESSC to describe the planet as

an interconnected whole that required interdisciplinary study. In retrospect, the ESSC’s work

contained two enmeshed components: promoting the study and understanding of the Earth as a

system, and advocating for a specific program—Earth system science—by which this could be done. The specific program placed a heavy reliance on satellite data and confronted the

“challenge” of developing a computational model of global change for shorter term Earth processes. The efforts that the ESSC undertook to promote ESS help explain why the specific phrase “Earth system” became the way that Earth scientists describe the planet as an interconnected system, instead of, say, the “global system” or the “coupled land-ocean- atmosphere system.” The ESSC put great effort into promoting its specific ESS research program and though that largely failed, they inadvertently provided the phrase that filled a semantic void in the Earth sciences. After 1986, the “Earth system” was the term that Earth scientists used almost exclusively to refer to the Earth as a scientific object.

The AGU case study illustrates how this happened. AGU members worried that the ESS project proposal by the ESSC marginalized the long timescales of geological and geophysical change by emphasizing the shorter timescales of atmospheric, oceanographic, and ecological change. They responded to this by arguing that ESS was not comprehensive enough to provide a full understanding of the Earth system. Many members maintained that this work needed to go further and study (variously) the whole, total, or complete Earth, including processes at work in the Earth’s core, mantle and outer environs. In doing so, they repeatedly rejected the narrow definition of the Earth system prioritized by the ESSC. However, as the AGU’s interdisciplinary activities and its various committees and initiatives show, its members were susceptible to and

340

readily adopted the Earth system phrase as the preferred way to capture the idea that the Earth

was a system of interconnected subsystems. Even while decrying ESS being “incomplete,”

AGU documents reveal that its members had, almost immediately after being exposed to ESSC

ideas via conference presentations and reports, adopted the term “Earth system” as the

predominant way to describe an interconnected planet. Through its collaborations with the ESA

and AMS, interdisciplinary conferences (including the Chapman Conference on the Gaia

hypothesis), the Earth-as-a-System Committee, the Planet Earth Initiative, and the Planet Earth

Committee, the AGU uncritically and unreflectively adopted the ESSC’s way of looking at

planet Earth. The AGU’s activities serve as an example of how the Earth system concept spread

beyond the ESSC to other communities and institutions.

That the AGU—a group that emphatically disagreed with the ESSC on so many key points—could still adopt the Earth system concept as a useful way to refer to the planet as a scientific object points to the ease with which that concept could travel among Earth scientists.

John Krige and other contributors to How Knowledge Moves (2019) rightly emphasize

impediments to the movement of knowledge and ideas, and so highlight the work involved in

overcoming these obstacles to construct a network of like-minded practitioners approaching scientific research with broadly similar conceptions and practices.163 The sudden prevalence of

the Earth system concept in the Earth sciences after 1986 shows what can happen when these

efforts are successful.

163 Krige, ed., How Knowledge Moves.

Epilogue: Mostly Harmless

Unless one is committed, there is hesitancy, the chance to draw back, always ineffectiveness. Concerning all acts of initiative (and creation), there is one elementary truth, the ignorance of which kills countless ideas and splendid plans: that the moment one definitely commits oneself, then providence moves too.

All sorts of things occur to help one that would never otherwise have occurred. A whole stream of events issues from the decision, raising in one’s favor all manner of unforeseen incidents and meetings and material assistance, which no man could have dreamed would have come his way.

Whatever you do, or dream you can, begin it. Boldness has genius, power and magic in it. Begin it now.

Goethe (as quoted on the ESSC’s Closer View publication folder, 1988)1

This dissertation ends with the early migration of the “Earth system” beyond the confines of the

ESSC, as exemplified by the AGU case study. Given how the “Earth system” was used in the

late 1980s, it still carried its original heuristic meaning prototyped by the ESSC. The entire

Earth was something to be studied as a system with interconnecting subsystems and referred to

by the single phrase developed and promoted by the ESSC. Earth scientists from different

specialist disciplines had different understandings of the specific contours of the Earth system,

but the broader and more colloquial sense of the phrase—an interconnected planet—resonated

universally. After the ESSC’s Earth system science reports appeared in 1986 and 1988, the use

of the term “Earth system” exploded.

The AGU aggressively rejected the ESSC’s specific ESS research program but warmly

embraced the “Earth system” phrase and the idea that the Earth was a system. That the AGU so

quickly adopted this phrase (and that the concept readily spread more widely) points to two key

things. First, there was an unfilled (and often unstated) need in Earth science disciplines for an

1 Quoted in: Earth System Sciences Committee (ESSC), Earth System Science: A Program For Global Change (Washington, DC: NASA, 1988).

341 342 uncontroversial phrase that captured the idea that the Earth was being conceived of as a system of systems. Borrowing from Leo Marx, I call this apparent unfilled need a “semantic void.”

Second, the vague and uncontroversial character of the “Earth system” concept made it tenable as a “boundary object” for the Earth sciences, in Susan Leigh Star’s sense of the term. Other phrases that might have filled the void—like “Gaia” or “global habitat”—could not be boundary objects in the relevant sense because they were not appropriately vague or uncontroversial. Its vague and innocuous character facilitated the Earth system’s spread and allowed scientists to talk about an interconnected planetary environment without saying all that much about the planet, its environment, or its interconnections. The AGU case study shows that, as the organization struggled to offer its own alternative to ESS, it could use the Earth system concept promoted by the ESSC without scruple. Indeed, as suggested, the AGU used the “Earth system” even more vaguely than the ESSC had. While the ESSC always maintained that it was a programmatic decision to study the Earth as a system, AGU committees used the “Earth system” with less ontological discretion, implying that the Earth and the Earth system were simply one and the same thing. The Earth was simply a system of interconnected subsystems. With this move, the

Earth system was on its way to being reified or naturalized.

A detailed examination of the further spread and eventual naturalization of the Earth system concept after 1989 is beyond the scope of this dissertation, but suggests a fruitful area of further research. The transition was almost certainly expedited by the continued development and deployment of Earth observing satellites and increasingly complex Earth system computer models throughout the 1990s and 2000s. While NASA never enacted the ESSC’s specific ESS research program, it did obtain Congressional approval and appropriations in 1989 for its

Mission to Planet Earth with its EOS satellite platforms that began flying in the late 1990s,

343 though in admittedly diminished forms.2 In 1999, Terra became the first EOS satellite to launch, with instruments collecting data on the Earth’s clouds, ice, water, land surfaces, and radiation budget, as well as atmospheric carbon monoxide.3 NASA’s EOS satellites, along with those from other space-faring countries, continue to contribute to the construction of global datasets for

Earth and environmental science research, thus facilitating the study of the planet as an interconnected system.4 Beginning in the 1980s, climate modelers created “Earth system models” by adding elements like sea ice, vegetation, snow cover, and agricultural land to models previously limited to certain atmospheric and oceanic components. By the 1990s, modelers had increased their scope, incorporating major carbon cycle processes into Earth system models, with later models—notably the IGBP’s AIMES—attempting to incorporate human actions into the system.5 These trends all contributed to the Earth system concept’s gradual entrenchment and

2 By the early 2000s, Erik Conway notes that NASA was the largest funder of climate change research in the US and one of the largest funders of Earth science research more broadly. Erik M. Conway, “Bringing NASA Back to Earth: A Search for Relevance during the Cold War,” in Science and Technology in the Global Cold War, eds. Naomi Oreskes and John Krige (Cambridge, MA: MIT Press, 2014), 251, 264. Work by Conway, W. Henry Lambright, Eric Goldstein, and Richard Leshner and Thor Hogan all detail NASA’s expansive (and expensive) vision for its EOS satellites, the political scrutiny it received in the 1990s, and the need to redesign, rebaseline, and rescope these satellites into smaller, more affordable, less complicated platforms with fewer remote sensing instruments. See: Erik M. Conway, Atmospheric Science at NASA: A History (Baltimore: Johns Hopkins University Press, 2008); Edward S. Goldstein, “NASA’s Earth Science Program: The Bureaucratic Struggles of the Space Agency’s Mission to Planet Earth” (PhD dissertation, George Washington University, 2007); W. Henry Lambright, “Administrative Entrepreneurship and Space Technology: The Ups and Downs of ‘Mission to Planet Earth,’” Public Administration Review 54, no. 2 (Mar./Apr. 1994): 97–104; W. Henry Lambright, “The Political Construction of Space Satellite Technology,” Science, Technology, and Human Values 19, no. 1 (Winter 1994): 47–69; Richard B. Leshner, “The Evolution of the NASA Earth Observing System: A Case Study in Policy and Project Formulation” (PhD dissertation, George Washington University, 2007); Richard B. Leshner and Thor Hogan, The View From Space: NASA's Evolving Struggle to Understand Our Home Planet (Lawrence, KS: University Press of Kansas, 2019). 3 “Terra: The EOS Flagship,” NASA, accessed 13 Mar 2020, https://terra.nasa.gov/. 4 See: “Open Data, Services and Software Policies,” EOSDIS, NASA, accessed 13 Mar 2020, https://earthdata.nasa.gov/collaborate/open-data-services-and-software. The WMO maintains a list of all current and future satellites that provide meteorological or other Earth observations. See: “List of All Satellites,” WMO, accessed 13 Mar 2020, https://www.wmo-sat.info/oscar/satellites. 5 Paul N. Edwards, A Vast Machine: Computer Models, Climate Data, and the Politics of Global Warming (Cambridge, MA: MIT Press, 2010), 146, 418; Ola Uhrqvist, “One Model to Fit All? The Pursuit of Integrated Earth System Models in GAIM and AIMES,” Historical Social Research 40, no. 2 (2015): 271-97.

344

unconscious naturalization by scientists, a process that proceeded from its development and promotion in the mid-1980s to its near-ubiquity today.

The specific ESS research program envisioned by the ESSC generated friction among some Earth science communities in the 1980s and was never fully implemented. However,

“Earth system science” has experienced a recent resurgence. Admittedly, still relatively few universities have “Earth system science” departments.6 Despite this lack, Earth and

environmental science departments offer many programs, courses, and symposia on “Earth

system science,” treating the Earth as a system comprised of subsystems in the atmosphere,

oceans, land, and biota, and make use of both global datasets and computer models.7 There are some scientists—notably Timothy Lenton—who self-identify as “Earth system scientists.”8 In the 1990s, Lenton studied nutrient and oxygen balances in the oceans and atmosphere under

Andrew Watson, who was himself a PhD student of James Lovelock’s in the late 1970s and early

1980s.9 While acknowledging the role of the ESSC in laying out what he called Earth system

science, Lenton defines Earth system science—in his contribution on “Earth System Science” for

Oxford University Press’ A Very Short Introduction series—as being completely decoupled from

any specific program. For Lenton, ESS is not a distinct research program with a detailed

implementation strategy, but simply a broad and interdisciplinary field of research that, “seeks to

understand how the planet functions as a whole system.”10 Earth system science could

6 A notable example is the Earth System Science Department at Stanford University, which offers graduate programs in ESS. See: “Earth System Science,” accessed 24 Apr 2020, https://earth.stanford.edu/ess. 7 A quick scan of almost any university’s Earth science course offerings will readily reveal these kinds of courses. 8 Lenton is the Director of the Global Systems Institute in the Geography Department at the University of Exeter. One of his “broad research specialisms” is “Earth system science.” See: “Professor Timothy Lenton,” University of Exeter, accessed 24 Apr 2020, http://geography.exeter.ac.uk/staff/index.php?web_id=Timothy_Lenton. 9 Watson and Lovelock developed the Daisyworld model to show how biota could unintentionally function as a self- regulating system that maintained stable conditions. See: Andrew J. Watson and James E. Lovelock, “Biological Homeostasis of the Global Environment: The Parable of Daisyworld,” Tellus B: Chemical and Physical Meteorology 35, no. 4 (1983): 284-9. 10 Tim Lenton, Earth System Science: A Very Short Introduction (New York: Oxford University Press, 2016), 1.

345

experience a resurgence once it had lost its former specificity, once it had become as vague as

the Earth system concept. Separated from its original programmatic connotations, Earth system

science can now, like the Earth system, serve as a multifaceted boundary object that facilitates interdisciplinary communication.

The entrenchment and near ubiquity of the Earth system concept in the Earth sciences—

as indicated by usage trends in Science, Nature, and on Google’s ngram viewer—may be mostly

harmless and even largely beneficial. The Earth system might not, at first glance, have any

problematic connotations. Indeed, the view that there is an interconnected planet bolsters Earth

science research today, including that done by climate scientists. Nevertheless, it is important to

remember that the Earth system is not a natural object, if by “natural” one means part of the

fundamental, unchanging order. It is a heuristic device that became so entrenched that it now

appears natural, normal, and inevitable. As historian of science Sarah Dry remarks about global

knowledge in general, “The concept of global knowledge is a powerful one. It may be one that

we feel we need today, but this does not make it either neutral or natural.”11 The same can be

said for the Earth system. Part of the work of this dissertation is to show the historically

contingent early roots of the Earth system concept. With the subsequent entrenchment of the

concept comes an ontological shift whereby the world now has a new kind of thing in it, a new

way to categorize and conceive of the planet. As Susan Leigh Star and Geoffrey Bowker remind

us in Sorting Things Out (1999), categorization schemes often remain invisible but have both material and practical consequences.12 How humans order the world affects how they act in it,

and the kinds of possibilities that might or might not present themselves.

11 Sarah Dry, Waters of the World: The Story of the Scientists Who Unraveled the Mysteries of Our Oceans, Atmosphere, and Ice Sheets and Made the Planet Whole (Chicago: University of Chicago Press, 2019), 274. 12 Geoffrey C. Bowker and Susan Leigh Star, Sorting Things Out: Classification and Its Consequences (Cambridge, MA: MIT Press, 1999).

346

While useful, the Earth system “model” should not be mistaken for the physical world.13

This material thing we live on is not itself a system. In systems approaches, the system is simply the unit of analysis, the thing being studied, be it a particular microbe, a plant, a tree, a forest, a body of water, or the entire planet. The system is not separable from the physical world. But it is also not a natural or inevitable way to order the world. Labeling something a system does not carve the world at its joints. The system is demarcated by those conducting the study, in this case, Earth scientists. As such, it is a human imposition on the physical world, a simplification made in an effort to grapple with the world’s complexities, to “conquer” the “abundance” of that physical world (to borrow from Paul Feyerabend).14

As this dissertation has shown, the Earth system was a vague concept and its vagueness facilitated its adoption. Its vagueness results, in part, from the thoroughly ambiguous character of the word “system” itself.15 There is, of course, the straightforward Oxford English Dictionary definition of a system: “an organized or connected group of things.”16 That definition is fine as far as it goes, but it elides the issue of what gets labelled as a system and the implications of that labelling. At the most basic (perhaps crude) level, there are two major interpretations of

“system”: reductionistic or holistic. For some, “system” implies the reduction of something into

13 A model represents certain key features of the physical world, but is not interchangeable or isomorphic with the physical world. Climate model ethnographers like Mayanna Lahsen and Simon Shackley observe that modelers can become so immersed in what Paul Edwards’ calls the “tuning” of model parameterizations to produce more accurate models that modelers can forget that the model is not actually the world but simply a representation of certain physical processes and components. See: Paul N. Edwards, A Vast Machine: Computer Models, Climate Data, and the Politics of Global Warming (Cambridge, MA: MIT Press, 2010), 344; Myanna Lahsen, “Seductive Simulations? Uncertainty Distribution Around Climate Models,” Social Studies of Science 35, no. 6 (Dec. 2005): 895-922; Simon Shackley, “Epistemic Lifestyles in Climate Change Modeling,” in Changing the Atmosphere: Expert Knowledge and Environmental Governance, ed. C. A. Miller and P. N. Edwards (MIT Press, 2001), 107-34. 14 Paul Feyerabend, Conquest of Abundance: A Tale of Abstraction Versus the Richness of Being, ed. Bert Terpstra (Chicago: University of Chicago Press, 1999). 15 In just one example, historian Howard Brick notes the contradictory interpretations of “system” in 1960s American, as both indicative of constrictive rational management and order, and as a holistic marker of the ecological interconnectedness of all things. See: Howard Brick, Age of Contradiction: American Thought and Culture in the 1960s (Ithaca, NY: Cornell University Press, 1998), ch. 6. 16 “System,” Oxford English Dictionary, accessed 21 Dec 2019, https://www.oed.com/.

347

simplified components for either analytical expediency or as a reflection of a simplified

worldview that effaces nuance and diversity. In this interpretation, defining something as a

system is a way of delimiting something as “contained” and comprised of a certain number of

components that can be identified and studied, taken apart and reassembled. In this view, the

system might be amenable to being fully known and, therefore, controlled by rational

managers.17 For others, labelling something a “system” identifies that thing as something

complex with emergent properties that cannot be understood by studying the individual

components in isolation. Here, to use a somewhat tired phrase, the sum is greater than its parts.

A system, in this interpretation, always has the capacity to surprise and one should remain epistemically modest when confronted with this uncertainty.18

An introductory survey of systems analysis suggests that whether one identifies a systems

approach, “as reductionistic or on the contrary as holistic is always a question of one’s point of

view.”19 In this way, “system” is comparable to Ludwig Wittgenstein’s interpretation of the ambiguous “duck-rabbit.” One might initially see only a duck or only a rabbit, not recognizing the possibility of seeing either in the same image. Once one becomes aware of the intrinsic ambiguity wherein a different cognitive ordering renders a different perception of the visual image, one can report seeing the image “as a duck” or “as a rabbit” and can shift back and forth

17 For an example of this interpretation, see: Sabine Höhler, Spaceship Earth in the Environmentalism Age, 1960- 1990 (New York: Routledge, 2016). 18 Andrew Pickering explicates this multifaceted character of systems in his history of British cyberneticians. They did not design and build systems to exert “control” over them, if by control one means “domination.” Rather, these cyberneticians sought to design systems capable of responding creatively to unanticipated external stimuli and behaving in unpredictable ways. Pickering argues that what motivated them was the desire to discern how humans might, “get along in a world that was not enframable, that could not be subjected to human designs.” See: Andrew Pickering, The Cybernetic Brain: Sketches of Another Future (Chicago: University of Chicago Press, 2010), 31-33. 19 Dieter M. Imboden and Stefan Pfenninger, Introduction to Systems Analysis: Mathematically Modeling Natural Systems (New York: Springer, 2013), 3.

348 between these interpretations.20 The key to understanding “system” is recognizing that it has this kind of “duck-rabbit” quality.

“System” is ambiguous in much the same way that the duck-rabbit is ambiguous. On the one hand, the term could be understood holistically, which lends itself to an emergent and epistemically modest understanding of systematicity. On the other hand, it could be interpreted reductively, which lends itself to a sense of systematicity that implies notions of total knowledge and control. The latter interpretation does not necessarily follow from any particular use of the term system. Simply to use “system” is not necessarily to implicate oneself in a program of technocratic rational management and domination. However, its ambiguous character does not preclude this interpretation. It will always be possible to impute this view when the term

“system” is used. It is always possible that “system” may be invoked to mean a reductionistic

“closed system” that is amenable to total understanding and control. Arguably, how the term system is invoked says much more about the interpreter than the system itself. This is why

Clifford Siskin argues that rather than trying to define “system,” it should instead be studied in its different contexts to see how it is being used in particular situations.21

The Earth scientists in this story did not use “system” in a reductionistic way. They deployed the Earth system concept as a heuristic device, as a broad organizational framework for thinking about and studying the planet as an interconnected whole. However, since its original formulation and use, the Earth system concept has become completely naturalized in the Earth and environmental sciences. In this naturalized state the concept could potentially be used in

20 Ludwig Wittgenstein, Philosophical Investigations, trans. G.E.M. Anscombe (Oxford, UK: Basil Blackwell, 1958), 194-7. 21 Clifford Siskin, System: The Shaping of Modern Knowledge (Cambridge, MA: MIT Press, 2016), 1-2.

349

reductionistic ways. What are the implications of this “Earth system” ambiguity for one of the

most pressing problems facing humanity, climate change?

In 2019, scientists and governments around the world declared a “climate emergency.”

One news outlet called 2019 “the year of ‘climate emergency’ declarations” as hundreds of

national governments made the acknowledgement, joined by over a thousand local governments

and over 11,000 scientists from 153 countries.22 From September 2018 to September 2019,

Oxford University Press recorded a 10,769% increase in the occurrences of “climate

emergency.” Because of this dramatic trend, it chose the phrase as the 2019 “Oxford Word of

the Year.”23 One reason for this trend may be the IPCC’s Special Report on “Global Warming

of 1.5°C” (2018) above pre-industrial temperatures. This report strongly recommends limiting

warming to 1.5°C (rather than 2°C) by drastically reducing greenhouse gas emissions by 2030

and reaching net zero emissions by 2050. The report warns that the consequences of the extra

0.5°C are far more severe than a “mere” 0.5°C might suggest. While still an achievable goal, the report notes the significant challenges and the unprecedented actions that would be required.24

Hence an “emergency” is now upon us. It remains to be seen whether this new call will effect

robust and lasting changes or lapse into symbolic gestures that afford little in the way of concrete

solutions.25 What is undeniable is that these declarations are supported by overwhelming

empirical evidence (aided in no small part by NASA’s fleet of Earth observing satellites).

22 Justine Culma, “2019 was the Year of ‘Climate Emergency’ Declarations,” The Verge, 27 Dec 2019, accessed 6 Jan 2020, https://www.theverge.com/2019/12/27/21038949/climate-change-2019-emergency-declaration; William J. Ripple, et al., “World Scientists’ Warning of a Climate Emergency,” BioScience 70, no. 1 (Jan. 2020): 8-12. 23 “Word of the Year 2019,” Oxford University Press, accessed 6 Jan 2020, https://languages.oup.com/word-of-the- year/2019/. 24 Intergovernmental Panel on Climate Change (IPCC), Global Warming of 1.5°C: An IPCC Special Report on the impacts of global warming of 1.5°C above pre-industrial levels and related global greenhouse gas emission pathways, in the context of strengthening the global response to the threat of climate change, sustainable development, and efforts to eradicate poverty (IPCC, 2018). 25 For instance, the Canadian federal government passed a non-binding motion that declared a climate emergency on June 17, 2019 and then the next day approved the Trans Mountain pipeline expansion, a pipeline the federal

350

How should governments, scientists, and the general public respond to this emergency, particularly given current political inertia and the steady increase in anthropogenic greenhouse gas emissions? One response in particular is gaining attention and will likely continue to grow in relevance. This option—geoengineering—covers a wide variety of methods.26 It broadly refers to the intentional alteration of planetary conditions to address the problems associated with climate change. More specifically, it means either removing carbon dioxide and other greenhouse gases from the atmosphere (, CDR) or reducing Earth’s radiation budget by decreasing incoming solar radiation (solar radiation management, SRM).27

The IPCC’s 2018 Special Report presented four “illustrative model pathways” to limit warming to 1.5°C. They all included some element of CDR geoengineering, which strongly suggests that reaching this target is highly unlikely without some reliance on geoengineering.28

Geoengineering is gaining such importance that it will receive unprecedented attention in the

IPCC’s upcoming Sixth Assessment Report (AR6) to be released beginning in 2021. Not only will CDR and SRM be incorporated into the report’s mitigation section, but they will, for the first time, be included in the section reserved for the “Physical Science Basis” of climate change,

government purchased in 2018 from Kinder Morgan for $4.5 billion. Critics argue this will increase greenhouse gas emissions in Alberta’s oil sands, thus making it difficult to reconcile the action with the “climate emergency” declaration. See: “House of Commons Declares a Climate Emergency Ahead of Pipeline Decision,” CBC News, 18 Jan 2019, accessed 6 Jan 2020, https://www.cbc.ca/news/politics/climate-emergency-motion-1.5179802; John Paul Tasker, “Trudeau Cabinet Approves Trans Mountain Expansion Project,” CBC News, 18 Jan 2019, accessed 6 Jan 2020, https://www.cbc.ca/news/politics/tasker-trans-mountain-trudeau-cabinet-decision-1.5180269. 26 For histories of earlier human ideas and/or attempts to alter the planet’s conditions, albeit on much smaller scales, see: J.R. McNeil and Corinna R. Unger, eds., Environmental Histories of the Cold War (New York: Cambridge University Press, 2010); James Rodger Fleming, Fixing the Sky: The Checkered History of Weather and Climate Control (New York : Columbia University Press, 2010); Kristine C. Harper, Make It Rain: State Control of the Atmosphere in Twentieth-Century America (Chicago: University of Chicago Press, 2017). 27 SRM also stands for “solar radiation modification” or “sunlight reflection methods.” See: National Research Council, Climate Intervention: Carbon Dioxide Removal and Reliable Sequestration (Washington, DC: The National Academies Press, 2015); National Research Council, Climate Intervention: Reflecting Sunlight to Cool the Planet (Washington, DC: The National Academies Press, 2015). 28 The amount of CDR varies between the four pathways. See: IPCC, Global Warming of 1.5°C, 14.

351

in which climatic responses to CDR and SRM will be examined.29 The IPCC’s “vision paper”

for the AR6’s scope listed CDR and SRM geoengineering as one of eight “cross-cutting issues”

the IPCC must include.30

All this is to say that geoengineering will receive new attention in the next IPCC report,

which will presumably, in turn, result in heightened attention from scientists, engineers,

government officials, and the general public. The intention here is not to overview these options

or weigh in on their efficacy, necessity, or desirability. Rather, it is to point to a potential issue

that might emerge when geoengineering ideas mix with a certain interpretation of the Earth

system. Suggested geoengineering techniques run the gamut from the smaller scale and more

technically feasible to the far-reaching and much more speculative. CDR strategies are generally

perceived as less invasive and problematic. SRM techniques, on the other hand, often carry with

them high social and environmental risks, and involve potentially dramatic and unpredictable

global effects. The most prominent and arguably technically feasible SRM strategy is

stratospheric aerosol injection, whereby particles (often sulfur dioxide) are injected into the

stratosphere to reflect sunlight, mimicking the observable cooling effects associated with sulfur

emissions from volcanic eruptions.31 Potential problems abound. They include effects on plant and animal populations, enhanced stratospheric ozone depletion, altered rainfall patterns that

29 “Report of the Forty-Sixth Session of the IPCC, Montreal, Canada, 6-10 September 2017,” accessed 8 Jan 2020, https://www.ipcc.ch/site/assets/uploads/2018/04/final_report_p46.pdf. The IPCC now calls CDR “greenhouse gas removal” (GGR). 30 “IPCC Chair’s Vision Paper: AR6 Scoping Meeting, Addis Ababa, Ethiopia, 1-5 May 2017,” accessed 8 Jan 2020, https://www.ipcc.ch/site/assets/uploads/2018/11/AR6-Chair-Vision-Paper.pdf. The US Congress recently provided $4 million for geoengineering research to David Fahey, Director of the Chemical Sciences Division of NOAA's Earth System Research Laboratory, to study various SRM techniques including sulfur dioxide injection into the stratosphere. See: John Fialka, “US Geoengineering Research Gets a Lift With $4 Million From Congress,” Science Magazine, 23 Jan 2020, accessed 13 Mar 2020, https://www.sciencemag.org/news/2020/01/us- geoengineering-research-gets-lift-4-million-congress. 31 P.J. Crutzen, “Albedo Enhancement By Stratospheric Sulfur Injections: A Contribution to Resolve a Policy Dilemma?” Climatic Change 77 (2006): 211-9. This technique is sometimes called stratospheric aerosol albedo modification (SAAM) or simply solar geoengineering. See: NRC, Climate Intervention: Reflecting Sunlight.

352

could create “winner” and “loser” countries, regulatory questions regarding who gets to “set” the

temperature, and the persistent need to reduce greenhouse gas emissions even if this technique is

successfully used. Past these issues, there are the particularly pesky and ineffable Rumsfeldian

“unknown unknowns.”32

Many scientists readily recognize the uncertainties and risks from geoengineering

techniques like stratospheric aerosol injection.33 However, planetary intervention on this scale

might begin out of desperation. The IPCC warns of the dramatic climatic effects that result from

a 2°C increase. If we do not take action today, future generations may face such conditions and

feel compelled to adopt widespread, risky strategies like SRM to dampen further climatic change

effects. Harvard University applied physicist David Keith frames the issue as a moral obligation.

We owe to future generations, he argues, serious study (with an initial emphasis on study rather

than use) of stratospheric injection to provide options for those in the future who may need to use

it due in large part to our inaction.34 Indeed, it hardly seems fair or reasonable to prejudge the

future use of risky and uncertain mitigation strategies if current generations fail to adequately

address the crisis.

But such a massive geo-intervention might also be motivated by human hubris fueled by

a reductionistic understanding of the Earth system. The reasoning here might be that if the Earth

system is fully understood then it can be radically geoengineered. When Earth scientists use the

phrase to describe an interconnected planet, they generally do not use it in a way that would have

32 NRC, Climate Intervention: Reflecting Sunlight, 94-5; David Keith, A Case for (Cambridge, MA: MIT Press, 2013), ch. 3 and 5. 33 One of the most prominent recent scientists advocating for the study of stratospheric aerosol injection techniques is Harvard University’s applied physicist David Keith. See: Keith, A Case for Climate Engineering; Juan B. Moreno-Cruz and David W. Keith, “Climate Policy Under Uncertainty: A Case for Solar Geoengineering,” Climatic Change 121, no. 3 (Dec. 2013): 431-44; John A. Dykema, et al., “Stratospheric Controlled Perturbation Experiment: A Small-Scale Experiment to Improve Understanding of the Risks of Solar Geoengineering,” Philosophical Transactions of the Royal Society A: Mathematical, Physical, and Engineering Sciences 372 (2014): 1-21. 34 David Keith, “What Role For Solar Geoengineering in Climate Policy?” (Distinguished Lecture Series, Department of Civil & Mineral Engineering, University of Toronto, Toronto, ON, 8 Nov 2019).

353

this implication. As this dissertation shows, the ESSC certainly did not hold a reductive view of

the Earth. The ESSC recognized that the Earth was a complex entity that was studied as a

system for practical purposes as a way of managing that complexity. The Earth system came to

refer to something exceedingly complex in its interconnections that required interdisciplinary

collaboration to improve understanding. However, the ambiguous “duck-rabbit” character of

“system” means that it can be invoked in a more reductionistic way than the ESSC ever intended.

“System” is inherently ambiguous. Someone promoting geoengineering solutions to the current

climate emergence could use a reductionistic view of the Earth system to justify the action and

diminish concerns about risks. The Earth system could be invoked to suggest that the Earth is

well-enough known that it will respond in predictable ways to large-scale planetary geo-

interventions.

The concern is not that reductionistic interpretations of the Earth system are in fact being

used to support large-scale, uncertain, potentially risky geoengineering strategies.35 The point at

issue is that, since geoengineering now factors into climate change mitigation, presumably some

will begin to argue for these projects in the future. They might make this argument by drawing

on a reductionistic interpretation of the Earth system, whereby the Earth system is something

contained, something known, and something amenable to management and control. The risk of

35 Some individuals promote more technological solutions to current global problems, notably the signatories— which include countercultural icon Stewart Brand, environmental scientist Erle Ellis, and Harvard applied physicist and SRM researcher David Keith—of the “Ecomodernist Manifesto.” While the Manifesto does not explicitly use the Earth system phrase to support its techno-optimistic vision for solving environmental issues, it does present a largely reductionistic and one-sided understanding of potential technological solutions without the humility that might come from the historical recognition that there are often unintended and detrimental consequences for technologies. It has many of the hallmarks of presenting simple solutions to complex problems. The Manifesto argues that humans must reduce their environmental impacts and can do so by accelerating technological innovations (through heavy government support) that intensify farming, energy extraction, forestry, and human settlements in order to benefit from using, “natural ecosystem flows and services more efficiently.” The Manifesto calls this a “good Anthropocene,” arguing that technology can and should be used to “decouple” human welfare from environmental impacts. Other technological innovations might include things like increased reliance on nuclear energy or geoengineering strategies. See: John Asafu-Adjaye, et al., “An Ecomodernist Manifesto,” Apr 2015, accessed 12 Jan 2020, http://www.ecomodernism.org/.

354

making the Earth an object of scientific knowledge is that it might be taken to imply that the

Earth has been “tamed,” that scientists possess complete working knowledge of how its parts do

and will interact.

The Earth system does not indicate any particular “state” of planetary knowledge. It remains a heuristic tool that recognizes that planetary complexity requires holistic study transcending traditional disciplinary boundaries. Earth scientists continue to study planetary conditions and interconnected Earth processes. The use of the phrase “Earth system” should not be taken to suggest that knowledge of the Earth is full and final, that this knowledge is, by and large, “worked out.” The Earth system is merely a model for understanding Earth as an interconnected planet requiring ongoing interdisciplinary research. A model of the Earth should not be mistaken for the physical world. The Earth system concept should not be used as part of a justification for confidently pursuing planetary interventions like SRM. Scientists might provide model-based arguments about the possible outcomes of geoengineering, but actual geoengineering intervention would be an experiment run in real time on the real planet. Given the Earth’s complex character, actually running the experiment would be the only way to know fully and finally what the outcomes will be. Ultimately, Earth science cannot settle the question or end debates about whether the risks of intervening outweigh the risks of not doing so. This should be a broader social and political debate and decision. De-naturalizing the Earth system concept is one way to ensure that it cannot be used to end debates that might soon confront us.

Bibliography

ARCHIVAL SOURCES

National Center for Atmospheric Research Archives, Boulder, CO.

National Aeronautics and Space Administration Historical Reference Collection (HRC), Washington, DC.

National Archives and Records Administration, College Park, MD.

PUBLISHED SOURCES

Abbate, Janet. Inventing the Internet. Cambridge, MA: MIT Press, 1999.

Adas, Michael. Machines as the Measure of Men: Science, Technology, and Ideologies of Western Dominance. Ithaca, NY: Cornell University Press, 1989.

“AGU Chapman Conferences.” EOS 68, No. 17 (28 Apr. 1987): 489.

Alexander, Joseph K. Science Advice to NASA: Conflict, Consensus, Partnership, Leadership. Washington, DC: NASA, 2017.

American Psychiatric Association. Diagnostic and Statistical Manual of Mental Disorders: DSM-5. Arlington, VA: American Psychiatric Association, 2013.

“Appointment of 14 Members of the National Commission on Space, and Designation of the Chairman and Vice Chairman.” Ronald Reagan Presidential Library. Accessed 27 Oct 2019. https://www.reaganlibrary.gov/research/speeches/32985a.

Arabatzis, Theodore. Representing Electrons: A Biographical Approach to Theoretical Entities. Chicago: University of Chicago Press, 2006.

Aronova, Elena. “Geophysical Datascapes of the Cold War: Politics and Practices of the World Data Centers in the 1950s and 1960s.” Osiris 32 (2017): 307-27.

Aronova, Elena, et al. “Introduction: Historicizing Big Data.” Osiris 32 (2017): 1-17.

Asafu-Adjaye, John et al. “An Ecomodernist Manifesto.” April 2015. Accessed 12 Jan 2020. http://www.ecomodernism.org/.

Ashley, Richard. “The Eye of Power: The Politics of World Modeling.” International Organization 37, No. 3 (Summer 1983): 495-535.

355 356

Barkley, Anne E., et al. “African Biomass Burning is a Substantial Source of Phosphorus Deposition to the Amazon, Tropical Atlantic Ocean, and Southern Ocean.” Proceedings of the National Academy of Sciences 116, No. 33 (13 Aug. 2019): 16216-21.

Barron, Eric J. Interviewed by Rebecca Wright, Tallahassee, FL, 1 Jul 2010. NASA Johnson Space Center, “Earth System Science at 20 Oral History Project.” Accessed 27 Oct 2019. https://historycollection.jsc.nasa.gov/JSCHistoryPortal/history/oral_histories/NASA_HQ/ ESS/BarronE/BarronE_7-1-10.pdf.

Barry, Roger G., and Richard J. Chorley. Atmosphere, Weather, and Climate, First Edition. London: Methuen & Co., 1968.

Beggs, James M. “Ambassador Beggs’ Statement, Aug. 10, 1982.” Department of State Bulletin 83, No. 2071 (Feb. 1983): 69-70.

Benson, Etienne S. “Re-Situating Fieldwork and Re-Narrating Disciplinary History in Global Mega-Geomorphology.” Studies in History and Philosophy of Science Part A 70 (2018): 28-37.

Bertalanffy, Ludwig von. General System Theory: Foundations, Development, Applications. New York: George Braziller, 1968.

Bertalanffy, Ludwig von. “An Outline of General System Theory.” The British Journal for the Philosophy of Science 1, No. 2 (Aug. 1950): 135-65.

Borges, Jorge Luis. “The Analytical Language of John Wilkins.” In Other Inquisitions, 1937- 1952, translated by Ruth L.C. Sims, 103. Austin, TX: University of Texas Press, 2000.

Bowker, Geoffrey C., and Susan Leigh Star. Sorting Things Out: Classification and Its Consequences. Cambridge, MA: MIT Press, 1999.

Bowler, Peter J. The Fontana History of the Environmental Sciences. London: Fontana Press, 1992.

Brennan, James R. “The Cold War Battle Over Global News in East Africa: Decolonization, the Free Flow of Information, and the Media Business, 1960-1980.” Journal of Global History 10 (2015): 333-56.

Bretherton, Francis P. “Earth System Science and Remote Sensing.” Proceedings of the Institute of Electrical and Electronics Engineers 73, No. 6 (Jun. 1985): 1118-27.

Bretherton, Francis P. “Ocean Climate Modeling.” Progress in Oceanography 11, No. 2 (1982): 93-129.

Bretherton, F.P., and M. Karweit. “Mid-Ocean Mesoscale Modeling.” In Numerical Models of Ocean Circulation: Proceedings of a Symposium Held at Durham, New Hampshire, October 17-20, 1972, 237-49. Washington, DC: National Academy of Sciences, 1975.

357

Brick, Howard. Age of Contradiction: American Thought and Culture in the 1960s. Ithaca, NY: Cornell University Press, 1998.

Burgess, Robert L. “The Ecological Society of America: Historical Data and Some Preliminary Analyses.” Ecological Society of America. Accessed 11 Jun 2019. https://www.esa.org/history/documents/BurgessHistory.pdf.

Burke, Kevin. “Planet Earth and the New Geoscience.” EOS 67, No. 4 (28 Jan. 1986): 42-3.

Butler, Dixon M. Interviewed by Rebecca Wright, Washington, DC, 25 Jun 2009. NASA Johnson Space Center “Earth System Science at 20 Oral History Project.” Accessed 10 Oct 2019. https://historycollection.jsc.nasa.gov/JSCHistoryPortal/history/oral_histories/NASA_HQ/ ESS/ButlerDM/ButlerDM_6-25-09.pdf.

Butler, Dixon M. Interviewed by Rebecca Wright, Washington, DC, 3 Jun 2010. NASA Johnson Space Center “Earth System Science at 20 Oral History Project.” Accessed 10 Oct 2019, https://historycollection.jsc.nasa.gov/JSCHistoryPortal/history/oral_histories/NASA_HQ/ ESS/ButlerDM/ButlerDM_6-3-10.pdf.

Camprubí, Lino. “The Invention of the Global Environment.” Historical Studies in the Natural Sciences 46, No. 2 (2016): 243-51.

Camprubí, Lino, and Philipp Lehmann. “The Scales of Experience: Introduction to the Special Issue Experiencing the global environment.” Studies in History and Philosophy of Science Part A 70 (2018): 1-5.

Cannon, Susan Faye. Science in Culture: The Early Victorian Period. New York: Science History Publications, 1978.

Carnap, Rudolph. “The Elimination of Metaphysics Through Logical Analysis of Language.” In Logical Positivism, edited by A.J. Ayer, 60-81. New York: The Free Press, 1959.

“Chapman Conferences: Guideline for Convenors.” EOS 56, No. 11 (Nov. 1975): 836-7.

Chen, H., et al. “Enhancing the ENSO Predictability Beyond the Spring Barrier.” Scientific Reports 10, No. 984 (2020): 1-12.

Cirac-Claveras, Gemma. “Satellites for What? Creating User Communities for Space-based Data in France: The Case from LERTS to CESBIO.” Technology and Culture 59, No. 2 (Apr. 2018): 203-25.

Clark, Bruce. “Gaia is not an Organism: Scenes from the Early Collaboration Between Lynn Margulis and James Lovelock.” In Lynn Margulis: The Life and Legacy of a Scientific Rebel, edited by Dorion Sagan, 32-43. White River Junction, VT: Chelsea Green Publishing, 2012.

358

Cloud, John. “Imagining the World in a Barrel: CORONA and the Clandestine Convergence of the Earth Sciences.” Social Studies of Science 31, No. 2 (Apr. 2001): 231-51.

Coen, Deborah R. “Big is a Thing of the Past: Climate Change and Methodology in the History of Ideas.” Journal of the History of Ideas 77, No. 2 (Apr. 2016): 305-21.

Coen, Deborah R. Climate in Motion: Science, Empire, and the Problem of Scale. Chicago: University of Chicago Press, 2018.

Committee on Earth Sciences, National Research Council. Our Changing Planet: A U.S. Strategy for Global Change Research. Washington, DC: National Academies Press, 1988.

Committee on Earth Sciences, Space Science Board, National Research Council. A Strategy for Earth Science from Space in the 1980s, Part I: Solid Earth and Oceans. Washington, DC: National Academy Press, 1982.

Committee on Earth Sciences, Space Science Board, National Research Council. A Strategy for Earth Science from Space in the 1980s, Part II: Atmosphere and Interactions With the Solid Earth, Oceans, and Biota. Washington, DC: National Academy Press, 1985.

Conway, Erik M. Atmospheric Science at NASA: A History. Baltimore, MD: Johns Hopkins University Press, 2008.

Conway, Erik M. “Bringing NASA Back to Earth: A Search for Relevance during the Cold War.” In Science and Technology in the Global Cold War, edited by Naomi Oreskes and John Krige, 251-72. Cambridge, MA: MIT Press, 2014.

Conway, Erik M. “Drowning in Data: Satellite Oceanography and Information Overload in the Earth Sciences.” Historical Studies in the Physical and Biological Sciences 37, No. 1 (2006): 127-51.

Conway, Erik M. “Planetary Science and the ‘Discovery’ of Global Warming.” In Exploring the Solar System: The History and Science of Planetary Exploration, edited by Roger D. Launius, 183-202. New York: Palgrave Macmillan, 2013.

Cosgrove, Denis. Apollo’s Eye: A Cartographic Genealogy of the Earth in the Western Imagination. Baltimore, MD: Johns Hopkins Press, 2001.

Cosgrove, Denis. “Contested Global Visions: One-World, Whole-Earth, and the Apollo Space Photographs.” Annals of the Association of American Geographers 84, No. 2 (1994): 270-94.

Crutzen, P.J. “Albedo Enhancement By Stratospheric Sulfur Injections: A Contribution to Resolve a Policy Dilemma?” Climatic Change 77 (2006): 211-9.

Crutzen, Paul J. “Geology of Mankind.” Nature 415, No. 3 (3 Jan. 2002): 23.

359

Crutzen, Paul J., and Eugene F. Stoermer, “The ‘Anthropocene.’” IGBP Newsletter 41 (May 2000): 17-8.

Culma, Justine. “2019 was the Year of ‘Climate Emergency’ Declarations.” The Verge, 27 Dec 2019. Accessed 6 Jan 2020. https://www.theverge.com/2019/12/27/21038949/climate- change-2019-emergency-declaration.

Daston, Lorraine, editor. Biographies of Scientific Objects. Chicago: University of Chicago Press, 2000.

Daston, Lorraine. “Historical Epistemology.” In Questions of Evidence: Proof, Practice, and Persuasion across the Disciplines, edited by James Chandler, Arnold I. Davidson, and Harry Harootunian, 282-9. Chicago: University of Chicago Press, 1994.

Daston, Lorraine. “Introduction: The Coming into Being of Scientific Objects.” In Biography of Scientific Objects, edited by Lorraine Daston, 1-14. Chicago: University of Chicago Press, 2000.

Daston, Lorraine. “Preface.” In Biographies of Scientific Objects, edited by Lorraine Daston, ix. Chicago: University of Chicago Press, 2000.

Daston, Lorraine, and Peter Galison. Objectivity. New York: Zone Books, 2007.

Dear, Peter. Discipline and Experience: The Mathematical Way in the Scientific Revolution. Chicago: University of Chicago Press, 1995.

DeSaussure, Hamilton. “Remote Sensing by Satellite: What Future for an International Regime?” The American Journal of International Law 71, No. 4 (Oct. 1977): 707-24.

Dessler, Andrew E. Introduction to Modern Climate Change. New York: Cambridge University Press, 2012.

Dettelbach, Michael. “The Face of Nature: Precise Measurement, Mapping, and Sensibility in the Work of Alexander von Humboldt.” Studies in History and Philosophy of Science Part C 30, No. 4 (1999): 473-504.

DeVorkin, David H. Science with a Vengeance: How the Military Created the US Space Sciences after World War II. New York: Springer-Verlag, 1992.

Doel, Ronald E. “American Geophysical Union.” In History of Science in the United States, edited by Marc Rothenberg, 30-1. New York: Garland Publishing Inc., 2001.

Doel, Ronald E. “Constituting the Postwar Earth Sciences: The Military’s Influence on the Environmental Sciences in the USA after 1945.” Social Studies of Science 35, No. 5 (Oct. 2003): 635-66.

360

Doel, Ronald E. “Scientists as Policymakers, Advisors, and Intelligence Agents: Linking Contemporary Diplomatic History with the History of Contemporary Science.” In The Historiography of Contemporary Science and Technology, edited by Thomas Söderqvist, 215-44. Amsterdam: Harwood Academic, 1997.

Doel, Ronald E. Solar System Astronomy in America: Communities, Patronage, and Interdisciplinary Research, 1920–1960. New York: Cambridge University Press, 1996.

Dry, Sarah. Waters of the World: The Story of the Scientists Who Unraveled the Mysteries of Our Oceans, Atmosphere, and Ice Sheets and Made the Planet Whole. Chicago: University of Chicago Press, 2019.

Dutreuil, Sébastien. “Gaïa: Hypothèse, Programme de Recherche Pour le Système Terre ou Philosophie de la Nature?” PhD diss., University of Paris 1 Pantheon-Sorbonne, 2016.

Dykema, John A., et al. “Stratospheric Controlled Perturbation Experiment: A Small-Scale Experiment to Improve Understanding of the Risks of Solar Geoengineering.” Philosophical Transactions of the Royal Society A: Mathematical, Physical, and Engineering Sciences 372 (2014): 1-21.

Earth System Sciences Committee (ESSC). Earth System Science: A Program For Global Change: A Closer View. Washington, DC: NASA, 1988.

Earth System Sciences Committee (ESSC). Earth System Science: A Program For Global Change: Overview. Washington, DC: NASA, 1986.

Easterbrook, Steve. “Climate Models as Boundary Objects.” Lecture, Issues in the Theoretical Foundations of Climate Science: Scientific and Philosophical Perspectives, Toronto, CA, 15 November 2018.

Edelson, Burton I. “Mission to Planet Earth.” Science 227 (25 Jan. 1985): 367.

Edelson, Edward. “Laying the Foundation.” Mosaic 19, No. 3/4 (Fall/Winter 1988): 4-11.

Edgerton Jr., Samuel Y. “Galileo, Florentine ‘Disegno,’ and the ‘Strange Spottednesse’ of the Moon.” Art Journal 44, No. 3 (1984): 225-32.

Edwards, Paul N. The Closed World: Computers and the Politics of Discourse in Cold War America. Cambridge, MA: MIT Press, 1996.

Edwards, Paul N. “Knowledge Infrastructures for the Anthropocene.” The Anthropocene Review 4, No. 1 (2017): 34-43.

Edwards, Paul N. “Representing the Global Atmosphere: Computer Models, Data, and Knowledge About Climate Change.” In Changing the Atmosphere: Expert Knowledge and Environmental Governance, edited by Clark A. Miller, Paul N. Edwards, 31-66. Cambridge, MA: MIT Press, 2001.

361

Edwards, Paul N. A Vast Machine: Computer Models, Climate Data, and the Politics of Global Warming. Cambridge, MA: MIT Press, 2010.

Edwards, Paul N. “The World in a Machine: Origins and Impacts of Early Computerized Global Systems Models.” In Systems, Experts, and Computers: The Systems Approach in Management and Engineering, World War II and After, edited by Agatha C. Hughes and Thomas P. Hughes, 221-53. Cambridge, MA: MIT Press, 2000.

Elichirigoity, Fernando. Planet Management: Limits To Growth, Computer Simulation, and the Emergence of Global Spaces. Evanston, IL : Northwestern University Press, 1999.

Escobar, Arturo. Encountering Development: The Making and Unmaking of the Third World. Princeton, NJ: Princeton University Press, 2011 [1995].

Escobedo, Tricia. “When a National Disaster Unfolded Live in 1986.” CNN. 31 March 2016. Accessed 9 October 2018. https://www.cnn.com/2016/03/31/us/80s-cnn-challenger- coverage/index.html.

Evans, D.L., et al. “Variations in the Earth's Orbit: Pacemaker of the Ice Ages?” Science 194, No. 4270 (10 Dec. 1976): 1121-1132.

“Executive Order 12545 -- National Commission on Space.” Ronald Reagan Presidential Library. Accessed 27 Oct 2019. https://www.reaganlibrary.gov/research/speeches/11486b.

Feyerabend, Paul. Conquest of Abundance: A Tale of Abstraction Versus the Richness of Being, edited by Bert Terpstra. Chicago: University of Chicago Press, 1999.

Fialka, John. “US Geoengineering Research Gets a Lift With $4 Million From Congress.” Science Magazine. 23 Jan 2020. Accessed 13 Mar 2020. https://www.sciencemag.org/news/2020/01/us-geoengineering-research-gets-lift-4- million-congress.

Fisher, Arthur. “One Model to Fit All.” Mosaic, 19, No. 3/4 (1988): 52-9.

Fisk, Leonard. Interviewed by Rebecca Wright, Ann Arbour, MI, 8 Sep 2010. NASA Johnson Space Center “Earth System Science at 20 Oral History Project.” Accessed 12 Aug 2019. https://historycollection.jsc.nasa.gov/JSCHistoryPortal/history/oral_histories/NASA_HQ/ Administrators/FiskLA/FiskLA_9-8-10.pdf.

Fleming, James Rodger. Fixing the Sky: The Checkered History of Weather and Climate Control. New York: Columbia University Press, 2010.

Fleming, James Rodger. Inventing Atmospheric Science: Bjerknes, Rossby, Wexler, and the Foundations of Modern Meteorology. Cambridge, MA: MIT Press, 2016.

362

Fleming, James Rodger, Vladimir Janković, and Deborah R. Coen, editors. Intimate Universality: Local and Global Themes in the History of Weather and Climate. Sagamore Beach, MA: Science History Publications/USA, 2006.

“Fooling Mother Nature: Lovelock Gives 25th Anniversary Lecture.” Staff Notes: National Center for Atmospheric Research 20, No. 39 (27 Sep. 1985): 4-5.

Foucault, Michel. “Discourse and Truth: The Problematization of Parrhesia (Six Lectures Given by Michel Foucault at the University of California at Berkeley, Oct-Nov. 1983).” Accessed 10 Dec 2019. https://www.cscd.osaka- u.ac.jp/user/rosaldo/On_Parrehesia_by_Foucault_1983.pdf.

Foucault, Michel. “What is Enlightenment?” In The Foucault Reader, edited by Paul Rabinow, 32-50. New York: Pantheon Books, 1984.

Galison, Peter. Image and Logic: A Material Culture of Microphysics. Chicago: University of Chicago Press, 1997.

Galison, Peter, and Emily Thompson, editors. The Architecture of Science. Cambridge, MA: MIT Press, 1999.

Ghilarov, Alexej M. “Vernadsky's Biosphere Concept: An Historical Perspective.” The Quarterly Review of Biology 70, No. 2 (Jun. 1995): 193-203.

Gilbreth, Frank B., and L.M. Gilbreth. Process Charts. New York: The American Society of Mechanical Engineers, 1921.

Gingras, Yves. “Naming Without Necessity: On the Genealogy and Uses of the Label ‘Historical Epistemology.’” Revue de Synthèse 131, No. 3 (2010): 439-54.

Global Atmospheric Research Programme (GARP) Joint Organising Committee (JOC). An Introduction to GARP. GARP Publication Series No. 1, October 1969.

Global Atmospheric Research Programme (GARP) Joint Organising Committee (JOC). The Physical Basis of Climate and Climate Modelling, Report of the International Study Conference in Stockholm, 29 July-10 August 1974. GARP Publication Series No. 16, April 1975.

Goldstein, Edward S. “NASA’s Earth Science Program: The Bureaucratic Struggles of the Space Agency’s Mission to Planet Earth.” PhD diss., George Washington University, 2007.

Good, Gregory A. “Sydney Chapman: Dynamo Behind the International Geophysical Year.” In Globalizing Polar Science: Reconsidering the International Polar and Geophysical Years, edited by Roger D. Launius, et al., 177-203. New York: Palgrave Macmillan, 2010.

363

Goodman, Nelson. Fact, Fiction, and Forecast. New York: The Bobbs-Merrill Company, 1965.

Gordin, Michael D. The Pseudoscience Wars: Immanuel Velikovsky and the Birth of the Modern Fringe. Chicago: University of Chicago Press, 2012.

Gribbin, John, and Mary Gribbin. James Lovelock: In Search of Gaia. Princeton, NJ: Princeton University Press, 2009.

Grinevald, Jacques. “Introduction: The Invisibility of the Vernadskian Revolution.” In The Biosphere, translated by David B. Langmuir, 20-32. New York: Copernicus, 1998.

Gwertzman, Bernard. “US is Quitting UNESCO, Affirms Backing for UN.” New York Times. 30 Dec 1983. Accessed 13 Mar 2020. https://www.nytimes.com/1983/12/30/world/us- is-quitting-unesco-affirms-backing-for-un.html.

Hacking, Ian. The Emergence of Probability: A Philosophical Study of Early Ideas About Probability, Induction and Statistical Inference. New York: Cambridge University Press, 1975.

Hacking, Ian. Historical Ontology. Cambridge, MA: Harvard University Press, 2002)

Hacking, Ian. “The Looping Effects of Human Kinds.” In Causal Cognition: An Interdisciplinary Approach, edited by D. Sperber, D. Premack, and A. Premack, 351-83. Oxford: Oxford University Press, 1995.

Hacking, Ian. Mad Travelers: Reflections On the Reality of Transient Mental Illnesses. Charlottesville, VA: University Press of Virginia, 1998.

Hacking, Ian. Rewriting the Soul: Multiple Personality and the Sciences of Memory. Princeton, NJ: Princeton University Press, 1995.

Haffner, Jeanne. The View From Above: The Science of Social Space. Cambridge, MA: MIT Press, 2013.

Hagen, Joel B. An Entangled Bank: The Origins of Ecosystem Ecology. New Brunswick, NJ: Rutgers University Press, 1992.

Hagen, Joel B. “Eugene Odum and the Homeostatic Ecosystem: The Resilience of an Idea.” In Traditions of Systems Theory: Major Figures and Contemporary Developments, edited by Darrell P. Arnold, 179-96. New York: Routledge, 2014.

Hamblin, Jacob Darwin. Arming Mother Nature: The Birth of Catastrophic Environmentalism. Toronto: Oxford University Press, 2013.

Hamblin, Jacob Darwin. Oceanographers and the Cold War: Disciples of Marine Science. Seattle: University of Washington Press, 2005.

364

Hammond, Debora. The Science of Synthesis: Exploring the Social Implications of General Systems Theory. Boulder, CO: University Press of Colorado, 2003.

Haraway, Donna J. Modest_Witness@Second_Millennium.FemaleMan©_Meets_OncoMouseTM: Feminism and Technosciences, second edition. New York: Routledge, 2018.

Haraway, Donna J. “Situated Knowledges: The Science Question in Feminism and the Privilege of Partial Perspective.” In Feminism and Science, edited by Evelyn Fox Keller and Helen E. Longino, 249-63. New York: Oxford University Press, 1996.

Harper, Dorothy. Eye in the Sky: Introduction to Remote Sensing, Second Edition. Montréal: Multiscience Publications, 1983.

Harper, Kristine C. Make It Rain: State Control of the Atmosphere in Twentieth-Century America. Chicago: University of Chicago Press, 2017.

Harper, Kristine C. Weather by the Numbers: The Genesis of Modern Meteorology. Cambridge, MA: MIT Press, 2008.

Hartree, Douglas. Calculating Instruments and Machine. Urbana, IL: University of Illinois Press, 1949.

Hayles, N. Katherine. How We Became Posthuman: Virtual Bodies in Cybernetics, Literature, and Informatics. Chicago: University of Chicago Press, 1999.

Hecht, Gabrielle. Being Nuclear: Africans and the Global Uranium Trade. Cambridge, MA: MIT Press, 2012.

Heims, Steve. The Cybernetic Group. Cambridge, MA: MIT Press, 1991.

Helmore, Edward. “Nike Hit With Lawsuit From Four Women Who Allege Gender Discrimination.” The Guardian. 10 Aug 2018. Accessed 19 Oct 2019. https://www.theguardian.com/business/2018/aug/10/nike-lawsuit-women-gender- discrimination.

Hempel, Carl. Philosophy of Natural Science. Upper Saddle River, NJ: Prentice Hall, 1966.

Henderson, Kathryn. On Line and On Paper: Visual Representations, Visual Culture, and Computer Graphics in Design Engineering. Cambridge, MA: MIT Press, 1999.

Hepler-Smith, Evan. “Grasping the Technical Image.” Historical Studies in the Natural Sciences 47, No. 1 (2017): 117-25.

Hitchcock, Dian R., and James E. Lovelock. “Life Detection By Atmospheric Analysis” Icarus 7 (1967): 149-59.

365

Höhler, Sabine. Spaceship Earth in the Environmentalism Age, 1960-1990. New York: Routledge, 2016.

“House of Commons Declares a Climate Emergency Ahead of Pipeline Decision.” CBC News. 18 Jan 2019. Accessed 6 Jan 2020. https://www.cbc.ca/news/politics/climate- emergency-motion-1.5179802.

Hughes, Agatha C., and Thomas P. Hughes, editors. Systems, Experts, and Computers: The Systems Approach in Management and Engineering, World War II and After. Cambridge, MA: MIT Press, 2000.

Hughes, Thomas P. “The Evolution of Large Technological Systems.” In The Social Construction of Technological Systems: New Directions in the Sociology and History of Technology, edited by Wiebe E. Bijker, et al., 51-82. Cambridge, MA: MIT Press, 1987.

Hughes, Thomas P. Rescuing Prometheus: Four Monumental Projects That Changed the Modern World. New York: Vintage Books, 2000.

Humboldt, Alexander von., and Aimé Bonpland. Essay on the Geography of Plants, edited by Stephen T. Jackson, translated by Sylvie Romanowski. Chicago: University of Chicago Press, 2008.

Hume, David. An Enquiry Concerning Human Understanding, edited by Tom L. Beauchamp. Toronto: Oxford University Press, 1999.

Hünemörder, Kai. “Environmental Crisis and Soft Politics: Détente and the Global Environment, 1968–1975.” In Environmental Histories of the Cold War, edited by J.R. McNeil and Corinna R. Unger, 257-76. New York: Cambridge University Press, 2010.

Hutchinson, G. Evelyn. “The Biosphere.” Scientific American 223, No. 3 (Sep. 1970): 45-53.

Hutchinson, G. Evelyn. “Circular Causal Systems in Ecology.” Annals: New York Academy of Sciences 50, No. 4 (1948): 221-46.

Imboden, Dieter M., and Stefan Pfenninger. Introduction to Systems Analysis: Mathematically Modeling Natural Systems. New York: Springer, 2013.

Intergovernmental Panel on Climate Change (IPCC). Global Warming of 1.5°C: An IPCC Special Report on the impacts of global warming of 1.5°C above pre-industrial levels and related global greenhouse gas emission pathways, in the context of strengthening the global response to the threat of climate change, sustainable development, and efforts to eradicate poverty. IPCC, 2018.

“IPCC Chair’s Vision Paper: AR6 Scoping Meeting, Addis Ababa, Ethiopia, 1-5 May 2017.” Accessed 8 Jan 2020. https://www.ipcc.ch/site/assets/uploads/2018/11/AR6-Chair- Vision-Paper.pdf.

366

Jackson, Stephen T. “Instruments Utilized in Developing the Tableau physique.” In Essay on the Geography of Plants, edited by Stephen T. Jackson, translated by Sylvie Romanowski, 221-6. Chicago: University of Chicago Press, 2008.

Jacobson, Michael C., et al. Earth System Science: From Biogeochemical Cycles to Global Change. New York: Elsevier Academic Press, 2006.

Jasanoff, Sheila. “Image and Consciousness: The Formation of Global Environmental Consciousness.” In Changing the Atmosphere: Expert Knowledge and Environmental Governance, edited by Clark A. Miller and Paul N. Edwards, 309-38. Cambridge, MA: MIT Press, 2001.

Jet Propulsion Laboratory (JPL), Global Change: Impacts on Habitability: A Scientific Basis for Assessment. [Washington, DC]: NASA, 1982.

Jones, Owen. “Woke-Washing: How Brands are Cashing in on the Culture Wars.” The Guardian. 23 May 2019. Accessed 30 Sep 2019. https://www.theguardian.com/media/2019/may/23/woke-washing-brands-cashing-in-on- culture-wars-owen-jones.

Kaiser, David. Drawing Theories Apart: The Dispersion of Feynman Diagrams in Postwar Physics. Chicago: University of Chicago Press, 2005.

Kaiser, David. How the Hippies Saved Physics: Science, Counterculture, and the Quantum Revival. New York: W.W. Norton, 2011.

Kauffman, Eric G. “The Gaia Controversy: AGU’s Chapman Conference.” EOS 69, No. 31 (2 Aug. 1988): 763-4.

Keith, David. A Case for Climate Engineering. Cambridge, MA: MIT Press, 2013.

Keith, David. “What Role For Solar Geoengineering in Climate Policy?” Distinguished Lecture Series, Department of Civil & Mineral Engineering, University of Toronto, Toronto, ON, 8 November 2019.

Kerr, Richard A. “Why Bigger Isn’t Better in Earth Observation.” Science 253, No. 5027 (27 Sep. 1991): 1481-2.

Kingsland, Sharon E. The Evolution of American Ecology, 1890-2000. Baltimore, MD: Johns Hopkins University Press, 2005.

Kirchner, James W. “The Gaia Hypotheses: Are They Testable? Are They Useful?” In Scientists on Gaia, edited by Stephen H. Schneider, Penelope J. Boston, 38-46. Cambridge, MA: MIT Press, 1991.

Klein, Naomi. No Logo: No Space, No Choice, No Jobs: Taking Aim at the Brand Bullies. London: Flamingo, 2000.

367

Kline, Ronald R. The Cybernetics Moment: Or Why We Call Our Age the Information Age. Baltimore, MD: Johns Hopkins Press, 2015.

Kranzberg, Melvin. “Technology and History: ‘Kranzberg's Laws.’” Technology and Culture 27, No. 3 (Jul. 1986): 544-60.

Krasner, Stephen D. Structural Conflict: The Third World Against Global Liberalism. Berkeley, CA: University of California Press, 1985.

Krige, John, editor. How Knowledge Moves: Writing the Transnational History of Science and Technology. Chicago: University of Chicago Press, 2019.

Krige, John. “Introduction: Writing the Transnational History of Science and Technology.” In How Knowledge Moves: Writing the Transnational History of Science and Technology, edited by John Krige, 1- 31. Chicago: University of Chicago Press, 2019.

Kuiper, Gerard P. “Preface.” In The Earth as a Planet, edited by Gerard P. Kuiper, v-vi. Chicago: University of Chicago Press, 1954.

Kwa, Chunglin. “Local Ecologies and Global Science: Discourses and Strategies of the International Geosphere-Biosphere Programme.” Social Studies of Science 35, No. 6 (Dec. 2005): 923-50.

Kwa, Chunglin. “Romantic and Baroque Conceptions of Complex Wholes in the Sciences.” In Complexities: Social Studies of Knowledge Practices, edited by John Law and Annemarie Mol, 23-52. Durham, NC: Duke University Press, 2002.

Lahsen, Myanna. “Seductive Simulations? Uncertainty Distribution Around Climate Models.” Social Studies of Science 35, No. 6 (Dec. 2005): 895-922.

Lambright, W. Henry. “Administrative Entrepreneurship and Space Technology: The Ups and Downs of ‘Mission to Planet Earth.’” Public Administration Review 54, No. 2 (Mar./Apr. 1994): 97-104.

Lambright, W. Henry. NASA and the Environment: The Case of Ozone Depletion. Washington, DC: NASA, 2005.

Lambright, W. Henry. “The Political Construction of Space Satellite Technology.” Science, Technology, and Human Values 19, No. 1 (Winter 1994): 47-69.

Landgrebe, David A. “Scanning the Issue: The Special Issue on Perceiving Earth’s Resources from Space.” Proceedings of the Institute of Electrical and Electronics Engineers 73, no. 6 (Jun. 1985): 947-9.

Latour, Bruno. Aramis, or the Love of Technology, translated by Catherine Porter. Cambridge, MA: Harvard University Press, 1996.

368

Latour, Bruno. “Drawing Things Together.” In Representation in Scientific Practice, edited by Michael Lynch, Steve Woolgar, 19-68. Cambridge, MA: MIT Press, 1990.

Latour, Bruno. “On the Partial Existence of Existing and Non existing Objects.” In Biographies of Scientific Objects, edited by Lorraine Daston, 247-69. Chicago: University of Chicago Press, 2000.

Latour, Bruno. Pandora’s Hope: Essays on the Reality of Science Studies. Cambridge, MA: Harvard University Press, 1999.

Latour, Bruno. The Pasteurization of France, translated by Alan Sheridan and John Law. Cambridge, MA: Harvard University Press, 1988.

Latour, Bruno. Politics of Nature: How to Bring the Sciences into Democracy, translated by Catherine Porter. Cambridge, MA: Harvard University Press, 2004.

Latour, Bruno. Science in Action: How to Follow Scientists and Engineers Through Society. Cambridge, MA: Harvard University Press, 1987.

Latour, Bruno. We Have Never Been Modern, translated by Catherine Porter. Cambridge, MA: Harvard University Press, 1993.

Launius, Roger D., James Rodger Fleming, and David H. DeVorkin, editors. Globalizing Polar Science: Reconsidering the International Polar and Geophysical Years. New York: Palgrave Macmillan, 2010.

Lenton, Tim. Earth System Science: A Very Short Introduction. New York: Oxford University Press, 2016.

Leshner, Richard B. “The Evolution of the NASA Earth Observing System: A Case Study in Policy and Project Formulation.” PhD diss., George Washington University, 2007.

Leshner, Richard B., and Thor Hogan. The View From Space: NASA's Evolving Struggle to Understand Our Home Planet. Lawrence, KS: University Press of Kansas, 2019.

Leslie, Stuart W. “‘A Different Kind of Beauty’: Scientific and Architectural Style in I. M. Pei's Mesa Laboratory and Louis Kahn's Salk Institute.” Historical Studies in the Natural Sciences 38, No. 2 (Spring 2008): 173-221.

Levy, H. The Universe of Science. London: Watts & Co., 1932.

Light, Jennifer. From Warfare to Welfare: Defense Intellectuals and Urban Problems in Cold War America. Baltimore, MD: Johns Hopkins Press, 2003.

Light, Jennifer. “When Women Were Computers.” Technology and Culture 40, No. 3 (Jul. 1999): 455-83.

369

Lombardozzi, Danica L., et al. “Temperature Acclimation of Photosynthesis and Respiration: A Key Uncertainty in the Carbon Cycle‐Climate Feedback.” Geophysical Research Letters 42 (2015): 8624-31.

Lovelock, J.E. Gaia: A New Look at Life on Earth. New York: Oxford University Press, 1987 [1979].

Lovelock, James. Homage to Gaia: The Life of an Independent Scientist. New York: Oxford University Press, 2000.

Lovelock, James. “Lecture Transcript.” National Bureau of Standards, UCAR/NCAR Oral History Collection, NCAR Archives. Accessed 12 Jun 2019. http://n2t.net/ark:/85065/d747485v.

Lovelock, J.E. “A Physical Basis For Life Detection Experiments.” Nature 207 (7 Aug 1965): 568-70.

Lovelock, James. The Revenge of Gaia: Why the Earth is Fighting Back - and How We Can Still Save Humanity. Toronto: Penguin Books, 2007.

Lovelock, James E. “A Sensitive Detector for Gas Chromatography.” Journal of Chromatography 1 (1958): 35-46.

Lovelock, James. The Vanishing Face of Gaia: A Final Warning. Toronto: Penguin Books, 2010.

Lovelock, J.E., R. J. Maggs, and R. A. Rasmussen. “Atmospheric Dimethyl Sulphide and the Natural Sulphur Cycle.” Nature 237 (23 Jun. 1972): 452-3.

Lovelock, James E., and Lynn Margulis. “Atmospheric Homeostasis By and For the Biosphere: The Gaia Hypothesis.” Tellus 26 (1974): 2-10.

Lovelock, James E., and Lynn Margulis. “Homeostatic Tendencies of the Earth’s Atmosphere.” Origins of Life 5 (1974): 93-103.

Lynch, Michael, and Steve Woolgar. “Introduction: Sociological Orientations to Representational Practice in Science.” In Representation in Scientific Practice, edited by Michael Lynch, Steve Woolgar, 1-18. Cambridge, MA: MIT Press, 1990.

MacBride Commission. Many Voices, One World: Towards a New, More Just and More Efficient World Information and Communication Order. New York: Unipub, 1980.

Mack, Pamela E. Viewing the Earth: The Social Construction of the Landsat Satellite System. Cambridge, MA: MIT Press, 1990.

370

Mack, Pamela E., and Ray A. Williamson. “Observing the Earth from Space.” In Exploring the Unknown: Selected Documents in the History of the US Civil Space Program, Vol. III: Using Space, edited by John M. Logsdon, 155-77. Washington, DC: NASA, [1998].

Maharaj, Ashok. “An Overview of NASA-India Relations.” In NASA and the World: Fifty Years of International Collaboration in Space, edited by John Krige, Angelina Long Callahan, and Ashok Maharaj, 211-34. New York: Palgrave Macmillan, 2013.

Maher, Neil M. Apollo in the Age of Aquarius. Cambridge, MA: Harvard University Press, 2017.

Maher, Neil M. “Bringing the Environment Back In: A Transnational History of Landsat.” In How Knowledge Moves: Writing the Transnational History of Science and Technology, edited by John Krige, 201-24. Chicago: University of Chicago Press, 2019.

Margulies, Lee. “TV Review: PBS ‘Earth’ Series Off to an Earthshaking Start.” Los Angeles Times. 22 Jan 1986. Accessed 23 Jun 2019. https://www.latimes.com/archives/la-xpm- 1986-01-22-ca-31688-story.html.

Margulis, Lynn. The Symbiotic Planet: A New Look at Evolution. London: Phoenix, 1998.

Margulis, Lynn, and J.E. Lovelock. “Biological Modulation of the Earth’s Atmosphere.” Icarus 21 (1974): 471-89.

Margulis, Lynn, et al. “Forward to the English Language Edition.” In The Biosphere, translated by David B. Langmuir, 14-9. New York: Copernicus, 1998.

Mark, Hans. “Observations of the Earth from Space: A Political Perspective.” 14 Sep 1983. Accessed 12 Feb 2018, https://historydms.hq.nasa.gov/.

Marsh, Peter. “Only the Rich Become Richer in Space.” New Scientist (19 Aug. 1982): 472-3.

Marx, Leo. “Technology: The Emergence of a Hazardous Concept.” Technology and Culture 51, No. 3 (Jul. 2010): 561-77.

Masco, Joseph. “The Age of Fallout.” History of the Present 5, No. 2 (Fall 2015): 137-68.

Masco, Joseph. “Bad Weather: On Planetary Crisis.” Social Studies of Science 40, No. 1 (Feb. 2010): 7-40.

Masco, Joseph. “Terraforming Planet Earth: The Age of Fallout.” In The Politics of Globality Since 1945: Assembling the Planet, edited by Rens van Munster and Casper Sylvest, 44- 70. New York: Routledge, 2016.

McCray, W. Patrick. “The Biggest Data of All: Making and Sharing a Digital Universe.” Osiris 32 (2017): 243-63.

371

McCray, Patrick. The Visioneers: How a Group of Elite Scientists Pursued Space Colonies, Nanotechnologies, and a Limitless Future. Princeton, NJ: Princeton University Press, 2013.

McElroy, John H., and Ray A. Williamson. “The Evolution of Earth Science Research from Space: NASA’s Earth Observing System.” In Exploring the Unknown: Selected Documents in the History of the US Civil Space Program: Volume VI: Space and Earth Science, edited by John M. Logsdon, et al., 441-73. Washington, DC: NASA, 2004.

McNeil, J.R., and Corinna R. Unger, editors. Environmental Histories of the Cold War. New York: Cambridge University Press, 2010.

McQuaid, Kim. “Selling the Space Age: NASA and Earth's Environment, 1958-1990.” Environment and History 12, No. 2 (May 2006): 127-63.

Meadows, Donella. Thinking in Systems: A Primer, edited by Diana Wright. White River Junction, VT: Chelsea Green Publishing, 2008.

Medina, Eden. Cybernetic Revolutionaries: Technology and Politics in Allende’s Chile. Cambridge, MA: MIT Press, 2011.

Messeri, Lisa. Placing Outer Space: An Earthly Ethnography of Other Worlds. Durham, NC: Duke University Press, 2016.

Messeri, Lisa. “The Problem with Pluto: Conflicting Cosmologies and the Classification of Planets.” Social Studies of Science 40, No. 2 (Apr. 2010): 187-214.

Mindell, David. Between Human and Machine: Feedback, Control, and Computing Before Cybernetics. Baltimore, MD: Johns Hopkins University Press, 2002.

Minger, Terrell J., editor. Greenhouse Glasnost: The Crisis of Global Warming. New York: Ecco Press, 1990.

Molina, Mario J., and F.S. Rowland. “Stratospheric Sink For Chlorofluoromethanes: Chlorine Atom Catalysed Destruction of Ozone.” Nature 249, No. 5460 (28 Jun 1974): 810-2.

Moor, Liz. The Rise of Brands. New York: Berg: 2007.

Moore, Berrien. Interviewed by Rebecca Wright, Norman, OK, 4 Apr 2011. NASA Johnson Space Center, “Earth System Science at 20 Oral History Project.” Accessed 10 Jun 2019. https://historycollection.jsc.nasa.gov/JSCHistoryPortal/history/oral_histories/NASA_HQ/ ESS/MooreB/MooreB_4-4-11.pdf.

“More About ‘Planet Earth.’” EOS 67, No. 4 (28 Jan. 1986): 43.

Moreno-Cruz, Juan B., and David W. Keith. “Climate Policy Under Uncertainty: A Case for Solar Geoengineering.” Climatic Change 121, No. 3 (Dec. 2013): 431-44.

372

Mulkay, Michael, and G. Nigel Gilbert. “Joking Apart: Some Recommendations Concerning the Analysis of Scientific Culture.” Social Studies of Science 12, No. 4 (1982): 585-613.

Munns, David P. Engineering the Environment: Phytotrons and the Quest for Climate Control in the Cold War. Pittsburgh: Pittsburgh University Press, 2017.

Murphy, Michelle. Sick Building Syndrome and the Problem of Uncertainty: Environmental Politics, Technoscience, and Women Workers. Durham, NC: Duke University Press, 2006.

NASA. Earth Observing System: Science Mission Requirements Working Group Report, Volume 1, Part 1 of 2. [Washington, DC]: NASA, [1985].

NASA. From Pattern to Process: The Strategy of the Earth Observing System, EOS Science Steering Committee Report Vol. II. [Washington, DC]: NASA, [1984].

NASA/NOSS Working Group. Needs, Opportunities and Strategies for a Long-Term Oceanic Sciences Satellite Program. Boulder, CO: National Center for Atmospheric Research, 1981.

National Research Council. Climate Intervention: Carbon Dioxide Removal and Reliable Sequestration. Washington, DC: The National Academies Press, 2015.

National Research Council. Climate Intervention: Reflecting Sunlight to Cool the Planet. Washington, DC: The National Academies Press, 2015.

National Research Council. An Implementation Plan for Priorities in Solar-system Space Physics. Washington, DC: National Academy Press, 1985.

Nicolson, Malcolm. “Alexander von Humboldt, Humboldtian Science and the Origins of the Study of Vegetation.” History of Science 25, No. 2 (Jun. 1987): 167-94.

Nye, Joseph. “Soft Power and American Foreign Policy.” Political Science Quarterly 119 (2004): 255-270.

Office of Technology Assessment (OTA). Remote Sensing and the Private Sector: Issues for Discussion. Washington, DC: GPO, 1984.

Office of Technology Assessment (OTA). UNISPACE ‘82: A Context for International Cooperation and Competition. Washington, DC: GPO, 1983.

Oreskes, Naomi, and John Krige, editors. Science and Technology in the Global Cold War. Cambridge, MA: MIT Press, 2014.

Pickering, Andrew. The Cybernetic Brain: Sketches of Another Future. Chicago: University of Chicago Press, 2010.

373

‘Planet Earth’: The Mystery With 100,000 Clues. [Washington, DC]: National Academy of Sciences, 1958.

Polavarapu, R.J., and G.L. Austin. “A Review of the GARP Atlantic Tropical Experiment (GATE).” Atmosphere-Ocean 17, No. 1 (1979): 1-13.

Poole, Robert. Earthrise: How Man First Saw the Earth. New Haven, CT: Yale University Press, 2008.

Popper, Karl R. Conjectures and Refutations: The Growth of Scientific Knowledge. New York: Basic Books, 1965.

Pratt, Mary Louise. Imperial Eyes: Travel Writing and Transculturation. New York: Routledge, 1992.

Priestley, Mark. A Science of Operations: Machines, Logic and the Invention of Programming. London: Springer, 2011.

Reinhold, Robert. “Summit of Sorts on Global Warming.” New York Times. 27 Aug 1989. Accessed 19 Oct 2019. https://www.nytimes.com/1989/08/27/us/summit-of-sorts-on- global-warming.html.

“Report of the Forty-Sixth Session of the IPCC, Montreal, Canada, 6-10 September 2017.” Accessed 8 Jan 2020. https://www.ipcc.ch/site/assets/uploads/2018/04/final_report_p46.pdf.

Ripple, William J., et al. “World Scientists’ Warning of a Climate Emergency.” BioScience 70, No. 1 (Jan. 2020): 8-12.

Romanowski, Sylvie. “Text of Humboldt’s Tableau physique.” In Essay on the Geography of Plants, edited by Stephen T. Jackson, translated by Sylvie Romanowski, 145-55. Chicago: University of Chicago Press, 2008.

Rothschild, Rachel Emma. Poisonous Skies: Acid Rain and the Globalization of Pollution. Chicago: University of Chicago Press, 2019.

Rudwick, Martin J.S. “The Emergence of a Visual Language for Geological Science, 1760- 1840.” History of Science 14, No. 3 (1976): 149-95.

Ruse, Michael. The Gaia Hypothesis: Science on a Pagan Planet. Chicago: University of Chicago Press, 2013.

Ryle, Gilbert. The Concept of Mind, Sixtieth Anniversary Edition. New York: Routledge, 2009.

Sachs, Aaron. “The Ultimate “Other”: Post-Colonialism and Alexander Von Humboldt's Ecological Relationship with Nature.” History and Theory 42, No. 4 (Dec. 2003): 111- 35.

374

Sagan, Carl, and George Mullen. “Earth and Mars: Evolution of Atmospheres and Surface Temperatures.” Science 177, No. 4043 (7 Jun. 1972): 52-6.

Sagan, Dorion, editor. Lynn Margulis: The Life and Legacy of a Scientific Rebel. White River Junction, VT: Chelsea Green Publishing, 2012.

Sargent, Daniel J. “North/South: The United States Responds to the New International Economic Order.” Humanity 6, No. 1 (Spring 2015): 201-16.

Schneider, Stephen H. “Climate Change and the World Predicament: A Case Study for Interdisciplinary Research.” Climatic Change 1 (1977): 21-43.

Schneider, Stephen H. “Editorial for the First Issue of Climatic Change.” Climatic Change 1, No. 1 (Mar. 1977): 3-4.

Schneider, Stephen H. Interviewed by Robert M. Chervin, 10-13 Jan 2002, Palo Alto, CA. Accessed 14 Jun 2019. http://n2t.net/ark:/85065/d72n50p5.

Selcer, Perrin. The Postwar Origins of the Global Environment: How the United Nations Built Spaceship Earth. New York: Columbia University Press, 2018.

Sellers, Christopher. Hazards of the Job: From Industrial Disease to Environmental Health Science. Chapel Hill, NC: University of North Carolina Press, 1997.

Shackley, Simon. “Epistemic Lifestyles in Climate Change Modeling.” In Changing the Atmosphere: Expert Knowledge and Environmental Governance, edited by Clark A. Miller, Paul N. Edwards, 107-34. Cambridge, MA: MIT Press, 2001.

Shilikhina, Ksenia. “‘Scientists Joke’: Evolution and Genres of Humour about Science and Scientists in Russia.” Styles of Communication 9, No. 2 (2017): 80-97.

Siddiqi, Asif A. Beyond Earth: A Chronicle of Deep Space Exploration, 1958-2016. Washington, DC: NASA, 2018.

Siskin, Clifford. System: The Shaping of Modern Knowledge. Cambridge, MA: MIT Press, 2016.

Slack, Nancy G. G. Evelyn Hutchinson and the Invention of Modern Ecology. New Haven, CT: Yale University Press, 2010.

Slotten, Hugh Richard. “Satellite Communications, Globalization, and the Cold War.” Technology and Culture 43, No. 2 (Apr. 2002): 315-50.

Smil, Vaclav. The Earth’s Biosphere: Evolution, Dynamics, and Change. Cambridge, MA: MIT Press, 2002.

375

Star, Susan Leigh. “The Structure of Ill-Structured Solutions: Boundary Objects and Heterogeneous Distributed Problem Solving.” In Readings in Distributed Artificial Intelligence, edited by Alan H. Bond and Les Gasser, 37-54. San Mateo, CA: Morgan Kaufmann Publishers, 1988.

Star, Susan Leigh. “This is Not a Boundary Object: Reflections on the Origin of a Concept.” Science, Technology, & Human Values 35, No. 5 (2010): 601-17.

Star, Susan Leigh, and James R. Griesemer. “Institutional Ecology, ‘Translations’ and Boundary Objects: Amateurs and Professionals in Berkeley's Museum of Vertebrate Zoology.” Social Studies of Science 19, No. 3 (Aug. 1989): 387-420.

Studies in History and Philosophy of Science Part A. Special Issue: Experiencing the Environment. 70 (Aug. 2018): 1-86.

Tansley, A.G. “The Use and Abuse of Vegetational Concepts and Terms.” Ecology 16, No. 3 (Jul. 1935): 283-307.

Task Group on Earth Sciences, Space Science Board, National Research Council. Mission to Planet Earth: Space Science in the Twenty-First Century: Imperatives for the Decades 1995 to 2015. Washington, DC: National Academy Press, 1988.

Tasker, John Paul. “Trudeau Cabinet Approves Trans Mountain Expansion Project.” CBC News. 18 Jan 2019. Accessed 6 Jan 2020. https://www.cbc.ca/news/politics/tasker- trans-mountain-trudeau-cabinet-decision-1.5180269.

Taylor, Peter J., and Ann S. Blum. “Ecosystems as Circuits: Diagrams and the Limits of Physical Analogies.” Biology and Philosophy 6 (1991): 275-94.

Tilford, Shelby G. “Global Habitability and Earth Remote Sensing.” Philosophical Transactions of the Royal Society of London. Series A, Mathematical and Physical Sciences 312, No. 1519 (Jul. 1984): 115-8.

Tilford, Shelby. Interview by Rebecca Wright, Washington, DC, 23 Jun 2009. NASA Johnson Space Center. “Earth System Science at 20 Oral History Project.” Accessed 5 Jun 2019. https://www.jsc.nasa.gov/history/oral_histories/NASA_HQ/ESS/TilfordSG/tilfordsg.htm.

Travis, G.D.L. “On the Construction of Creativity: The ‘Memory Transfer’ Phenomenon and the Importance of Being Earnest.” In The Social Process of Scientific Investigation, edited by Karin D. Knorr, et al., 165-193. Dordrecht: Reidel, 1981.

Tufte, Edward. The Visual Display of Quantitative Information, Second Edition. Cheshire, CT: Graphics Press, 2001.

Turchetti, Simone, and Peder Roberts, editors. The Surveillance Imperative: Geosciences During the Cold War and Beyond. New York: Palgrave Macmillan, 2014.

376

Uhrqvist, Ola. “One Model to Fit All? The Pursuit of Integrated Earth System Models in GAIM and AIMES.” Historical Social Research 40, No. 2 (2015): 271-97.

Uhrqvist, Ola. “Seeing and Knowing the Earth as a System: An Effective History of Global Environmental Change Research as Scientific and Political Practice.” PhD diss., Linköping University, 2014.

United Nations, Committee on the Peaceful Uses of Outer Space (COPUOS). Report of the Committee on the Peaceful Uses of Outer Space. New York: United Nations, 1968.

United Nations, General Assembly. Declaration on the Establishment of a New International Economic Order, A/RES/S-6/3201 (1 May 1974).

United Nations. Report of the Second United Nations Conference on the Exploration and Peaceful Uses of Outer Space, A/CONF.101/10. 31 Aug 1982.

United States Committee for the Global Atmospheric Research Program (GARP), National Research Council. Understanding Climatic Change: A Program for Action. Washington, DC: National Academy of Sciences, 1975.

United States House of Representatives, Committee on Foreign Affairs. Report on the Second U.N. Conference on the Peaceful Uses of Outer Space (UNISPACE 1982), August 9-21, 1982, 97th Cong., 2nd Sess. Washington, DC: GPO, 1983.

United States Senate. NASA’s Space Science Programs and the Mission to Planet Earth: Hearing before the Subcommittee on Science, Technology, and Space of the Committee on Commerce, Science, and Transportation, United States Senate, One Hundred Second Congress, First Session, April 24, 1991. Washington, DC: GPO, 1991.

“United States Withdrawal from UNESCO.” International Legal Materials 23, No. 1 (Jan. 1984): 221.

University Corporation for Atmospheric Research (UCAR). “Preliminary Plans for a National Institute for Atmospheric Research.” Accessed 27 Oct 2019. https://opensky.ucar.edu/islandora/object/archives:3054.

Vance, Tiffany C., and Ronald E. Doel. “Graphical Methods and Cold War Scientific Practice: The Stommel Diagram’s Intriguing Journey from the Physical to the Biological Environmental Sciences.” Historical Studies in the Natural Sciences 40, No. 1 (2010): 1- 47.

Van Munster, Rens, and Casper Sylvest. “Introduction.” In The Politics of Globality Since 1945: Assembling the Planet, edited by Rens van Munster and Casper Sylvest, 1-19. New York: Routledge, 2016.

Vernadsky, Vladimir I. The Biosphere, translated by David B. Langmuir. New York: Copernicus, 1998.

377

Vertesi, Janet. “Drawing as: Distinctions and Disambiguation in Digital Images of Mars.” In Representation in Scientific Practice Revisited, edited by Catelijne Coopmans, et al., 15- 35. Cambridge, MA: MIT Press, 2014.

Vertesi, Janet. Seeing Like a Rover: How Robots, Teams, and Images Craft Knowledge of Mars. Chicago: University of Chicago Press, 2015.

Wainwright, John. “Earth-System Science.” In A Companion to Environmental Geography, edited by Noel Castree, et al., 145-67. Malden, MA: Wiley-Blackwell, 2009.

Waldrop, M. Mitchell. “An Inquiry into the State of the Earth.” Science 226, No. 4670 (5 Oct. 1984): 33-5.

Watson, Andrew J., and James E. Lovelock. “Biological Homeostasis of the Global Environment: The Parable of Daisyworld.” Tellus B: Chemical and Physical Meteorology 35, no. 4 (1983): 284-9.

Weart, Spencer R. The Discovery of Global Warming. Cambridge, MA: Harvard University Press, 2003.

Wickland, Diane E. Interviewed by Rebecca Wright, Washington, DC, 26 Mar 2010. NASA Johnson Space Center “Earth System Science at 20 Oral History Project.” Accessed 27 Oct 2019. https://historycollection.jsc.nasa.gov/JSCHistoryPortal/history/oral_histories/NASA_HQ/ ESS/WicklandDE/WicklandDE_3-26-10.pdf.

Wiener, Norbert. Cybernetics: Or Control and Communication in the Animal and the Machine, Second Edition. Cambridge, MA: MIT Press, 1961.

Wiener, Norbert. The Human Use of Human Beings: Cybernetics and Society. New York: Doubleday, 1954.

Williams, Raymond. Culture and Society, 1780-1950. New York: Columbia University Press, 1983.

Williamson, Ray A. “The Landsat Legacy: Remote Sensing Policy and the Development of Commercial Remote Sensing.” Photogrammetric Engineering & Remote Sensing 63, No. 7 (Jul. 1997): 877-85.

Wittgenstein, Ludwig. Philosophical Investigations, translated by G.E.M. Anscombe. Oxford, UK: Basil Blackwell, 1958.

Wolfe, Audra J. Competing With the Soviets: Science, Technology, and the State in Cold War America. Baltimore, MD: The Johns Hopkins University Press, 2013.

Wolfe, Audra J. Freedom's Laboratory: The Cold War Struggle For the Soul of Science. Baltimore, MD: Johns Hopkins University Press, 2018.

378

Worster, Donald. Nature’s Economy: A History of Ecological Ideas, Second Edition. New York: Cambridge University Press, 1994.