MAKING GLOBAL WARMING GREEN: AND AMERICAN ENVIRONMENTALISM, 1957-1992

A DISSERTATION SUBMITTED TO THE DEPARTMENT OF HISTORY AND THE COMMITTEE ON GRADUATE STUDIES OF STANFORD UNIVERSITY IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

Joshua P. Howe July 2010

© 2010 by Joshua Proctor Howe. All Rights Reserved. Re-distributed by Stanford University under license with the author.

This work is licensed under a Creative Commons Attribution- Noncommercial 3.0 License. http://creativecommons.org/licenses/by-nc/3.0/us/

This dissertation is online at: http://purl.stanford.edu/cp892qc1059

ii I certify that I have read this dissertation and that, in my opinion, it is fully adequate in scope and quality as a dissertation for the degree of Doctor of Philosophy.

Richard White, Primary Adviser

I certify that I have read this dissertation and that, in my opinion, it is fully adequate in scope and quality as a dissertation for the degree of Doctor of Philosophy.

Robert Proctor

I certify that I have read this dissertation and that, in my opinion, it is fully adequate in scope and quality as a dissertation for the degree of Doctor of Philosophy.

Jessica Riskin

Approved for the Stanford University Committee on Graduate Studies. Patricia J. Gumport, Vice Provost Graduate Education

This signature page was generated electronically upon submission of this dissertation in electronic format. An original signed hard copy of the signature page is on file in University Archives.

iii Abstract Making Global Warming Green: Climate Change and American Environmentalism, 1957-1992 investigates how global climate change became a major issue in American environmental politics during the second half of the 20th Century. The dissertation focuses on the complex institutional, political, and professional relationships between scientists studying climate and America’s professional environmentalists during this tumultuous period. Throughout the early history of global warming, the geographical and chronological scales of climatic and atmospheric change transcended the existing legal and regulatory mechanisms of American environmental politics. Only scientists had the technology and expertise to recognize threats to these novel components of the environment, and their exclusive access all but forced them into a prominent role as environmental advocates. But scientists’ particular forms of advocacy often reflected the values and interests of their disciplines more than they did the middle-class quality-of-life concerns at the center of the mainstream American environmental movement. Scientists framed climate change in terms of development and natural resources, and they sought to influence elites in government and at international scientific organizations more than they worked to mobilize the American public. As American environmentalists began to take up the fight against global warming in the 1980s and ‘90s, they too found themselves trading in the language of scientific consensus and government-sponsored global solutions rather than the locally focused, middle-class consumer values originally at the heart of the movement. More than just a political history of global warming, this research presents a new history of American environmentalism that takes into account the science and politics of an issue that has emerged in the twenty-first century as one of our greatest challenges.

iv TABLE OF CONTENTS Abstract…………………………………………………….iv

Introduction………………………………………………...1

Chapter 1: The Cold War Roots of Global Warming……..19

Chapter 2: The Supersonic Transport: Scientists, Environmentalists, and the Atmospheric Environment……………………………………….67

Chapter 3: Systems Science, the Stockholm Conference, and the Making of the Global Environment……………………………………….112

Chapter 4: Climate, “The Environment,” and Scientific Activism in the 1970s………………………………………………166

Chapter 5: Scientists, Environmentalists, Democrats: The New Politics of Climate Change Under Ronald Reagan…………………….228

Chapter 6: Mechanisms of Change: Knowledge and Regulation in a Warming World……………………………………………...283

Conclusion………………………………………………...324

Bibliography………………………………………………338

v Introduction

The history of global warming is at once a romance and a tragedy. As a romance, it is a story of remarkable scientific and political achievement. Between 1957 and 1992, a relatively small community of scientists and environmentalists transformed a scientific curiosity at the fringes of the Cold War research system into the centerpiece of both

American and international environmentalism. Atmospheric scientists studying climate and climatic change built new institutions, established new disciplines, launched unprecedented cooperative international research initiatives, and ultimately created a new way of understanding the global atmosphere and humans’ relationship to it. Alongside

American environmentalists, these scientists introduced the problems of the global

atmosphere—and the threat of anthropogenic CO2-induced climate change in particular—into American politics and public life. They promoted and helped to shape an international framework for regulating emissions based on international cooperation and the best available scientific knowledge. Their success has led our society to begin proactively seeking solutions to global warming, and in this sense Making Global

Warming Green, much like Spencer Weart’s survey of the history of climate science in the 20th Century, The Discovery of Global Warming, tells an optimistic tale.1

But concomitant to the story of scientific and political success on climate change is a story of a larger political failure. Despite scientists’ and environmentalists’ remarkable efforts to study, popularize, and advocate for action on global warming in the

1 In narrating this story of success I diverge from Weart in my reliance on a mix of archival and published sources and in my persistent focus on the place of climate change in American environmentalism, but to the extent that Making Global Warming Green tells the story of the development of a scientific field—climate science—I, too, present that story as the qualified scientific triumph I believe that it is. Spencer Weart, The Discovery of Global Warming (Cambridge: Press, 2003), xvii.

1 20th century, these groups have after fifty years almost categorically failed to prevent the type of environmental change they have warned against for so long.2 Fossil fuel energy continues to support an expanding global economy, and greenhouse gas emissions consequently continue to expand as well. Environmentalists have certainly inspired gains in energy efficiency and land use changes around the world, but for the most part the climatic benefits of these gains are artifacts of more traditional local and regional efforts to control other environmental problems like common air pollution and deforestation.

Concentrated at about 315ppm in 1959, as of May 2010, CO2 sits at 392.24ppm. The rate of increase has risen from around 1ppm per year to closer to 2ppm per year.3 The problem itself is not getting better; it is getting worse.

How to reconcile these two stories? Within the limited historiography that exists, historians of global warming have tended to separate scientific success from political failure.4 This separation is in part an artifact of chronology. Climate scientists found

2 In a remarkably self-aware book, Red Sky at Morning: America and the Crisis of the Global Environment (New Haven: Yale University Press, 2004), James Gustave Speth takes the failure to produce action on global environmental problems as a jumping off point for an analysis of the global environmental politics in the last quarter of the 20th century. 3 For information on current CO2 concentrations, see the organization “CO2 Now,” http://co2now.org/; See also “Carbon Dioxide, Methane Rise Sharply in 2007,” NOAA, April 23, 2008, http://www.noaanews.noaa.gov/stories2008/20080423_methane.html 4 Critical histories of global warming are few and far between, but a few important works have laid the groundwork for the field. If there is a single, standard narrative of the history of the science of global warming, it is Weart’s The Discovery of Global Warming. The accompanying website, The Discovery of Global Warming: A Hypertext History of How Scientists Came to (Partly) Understand What People are Doing to Cause Climate Change, also provides a wealth of resources and discussions on interactions within the scientific community and between science and politics, the media, the environmental movement, the international community, and the public. The interpretive framework is very schematic, but the lack of nuance does not detract from its utility; indeed, it may facilitate it. James Roger Fleming offers perhaps the most critical and nuanced historical work on the science of climate, both in myriad articles on weather modification and in his Historical Perspectives on Climate Change (Oxford: Oxford University Press, 1998). Fleming is careful to note how the ideas about the concept of climate—its scale, its impact, and its permanence—have changed over time, although he, like other historians of science, has a tendency to focus on the details of “first references” and “discoveries,” some times to the detriment of his discussion of these events’ historical significance. Erik Conway’s Atmospheric Science at NASA: A History (Baltimore, MD: Johns Hopkins Press, 2008) attacks the history of atmospheric science from the perspective of a single but extremely important government agency, and though it does not aspire to be a history of global warming, it

2 their greatest successes—individual, institutional, and political—between 1957 and 1988; since 1988, their efforts to create political solutions to the problem of global warming have found strong opposition, and have largely failed. Whereas Weart lauds scientists for their discovery, investigation, and advocacy in the ‘60s, ‘70s, and ‘80s, scholars like

Naomi Oreskes and Eric Conway counterpose Weart’s work with a discussion of the villainy of the scientists, corporations, and government officials who have stood in the way of action on global warming in the ‘80s, ‘90s, and ‘2000s.5

The distinction between scientific success and political failure goes beyond simple chronology, however. Few serious scholars maintain the naïve assumption that science exists as an independent entity detached from politics—in fact, as historians and sociologists of science, they strive to understand the complex relationships between these two human endeavors. Still, Weart, Oreskes, and others lament what they frame as the

adds depth and detail to Weart’s survey. Most recently, Naomi Oreskes has teamed up with Conway to produce Merchants of Doubt: How a Handful of Scientists Obscured the Truth on Issues from Tobacco Smoke to Global Warming (New York: Bloomsbury, 2010), a book that investigates the nature of scientific skepticism not only on global warming, but also on tobacco, ozone, and other scientific controversies of the second half of the 20th Century. Chapters 6 and 7 focus specifically on global wrming. Weart and Oreskes write for mixed academic and popular audiences; in addition to their work there are a number of easily accessible journalistic histories of global warming that provide useful overviews of the development of the issue. For a very readable history of climate science that traces the field’s meteorological roots, see William K. Stevens, The Change in the Weather: People, Weather, and the Science of Climate (New York: Dell Publishing, 1999); for another popular history worth reading, see Gale E. Christianson, Greenhouse: the 200-year Story of Global Warming (New York: Penguin Books, 1999). For popular overview of the history itself, it is difficult to beat Tim Flannery’s The Weather Makers: How Man is Changing the Climate and What it Means for Life on (New York: Atlantic Monthly Press, 2005), although there are too many recent exposés to list here. 5 See Oreskes and Conway, “The Denial of Global Warming,” Chapter 6 in Merchants of Doubt, 169-215. See also Naomi Oreskes, Erik Conway, and Matthew Shindell, “From Chicken Little to Dr. Pangloss: William Nierenberg, Global Warming, and the Social Deconstruction of Scientific Knowledge,” Historical Studies in the Natural Sciences, vol. 38, no. 1 (Winter, 2008): 109-152; Mark Bowen, Censoring Science: Inside the Political Attack on Dr. James Hansen and the Truth of Global Warming (New York: Plume, 2008). Perhaps the most engaging history of the politics of global warming in the 1990s is Jeremy Leggett’s The Carbon War: Global Warming and the End of the Oil Era (New York: Routledge, 2001). For an example of a more polemical work, see James Hoggan with Richard Littlemore, Climate Cover-Up: The Crusade to Deny Global Warming (New York: Greystone Books, 2009). Much more nuanced is Michael Hulme’s Why We Disagree About Climate Change: Understanding Controversy, Inaction, and Opportunity (Cambridge: Cambridge University Press, 2009).

3 destructive politicization of climate science in the past three decades.6 Separating the story of scientists’ successes in studying global warming from their failed efforts to affect meaningful political change on the issue helps to reinforce this lament. It allows both historians and their living protagonists to lay the blame for the collective failure to take action on global warming solely at the feet of good scientists’ opponents.

It would be a mistake to overstate the case—Weart and Oreskes, for example, present nuanced and critical arguments about important aspects of the history of global warming, and their work gives shape to a historiography that is still very new. Moreover, as Oreskes and Conway chronicle in Merchants of Doubt, politically conservative politicians, institutions committed to free-market principles, skeptical and contrarian scientists, and officials from self-interested, energy-dependent governments have in the last twenty-five years consistently and actively stood in the way of real progress on global warming. There is plenty of blame to go around. But beyond Gus Speth’s remarkably honest and self-aware Red Sky at Morning, few scholars have investigated the liabilities inherent to the structures of science and advocacy developed in the second half of the 20th century by scientists and environmentalists themselves.7

The purpose of this dissertation is not to blame scientists for political inaction on global warming; global warming is certainly not their fault, nor are scientists necessarily responsible for the difficulties of transforming their knowledge into workable political solutions to a complex problem beyond their political control and often outside of their

6 See, for example, Ross Gelbspan, Boiling Point: How Politicians, Big Oil and Coal, Journalists, and Activists Have Fueled a Climate Crisis—And What We Can Do to Avert Disaster (New York: Basic Books, 2004) and The Heat is On: The Climate Crisis, the Cover-up, the Prescription (New York: Basic Books, 1998). For an insider’s account of how scientists watched their work take on a political life of its own, see Stephen H. Schneider, Science as a Contact Sport: Inside the Battle to Save Earth’s Climate (Washington, D.C.: National Geographic, 2009). 7 Speth, Red Sky at Morning.

4 professional milieu. Indeed, to the extent that critical histories have “good guys,” in my mind the scientists and environmentalists who have earnestly studied and fought to curb global warming are the good guys. But in combination with the scale and complexity of the problem itself and the powerful opposition to political action on the issue that arose in the 1980s, I contend that the very institutions and forms of advocacy scientists and environmentalists developed to study and mitigate global warming in the 20th century have contributed to the crisis we face today. These groups have identified and defined the problem, built the forums in which to attempt to solve it, and set the terms of debate.

And, to a large degree, their efforts have come up short. For advocates of action on global warming, Making Global Warming Green is thus fodder for self-reflection. It is an effort to understand first how scientists, environmentalists, and politicians created intellectual, institutional, and political frameworks for studying and eventually governing climate change, and second, why these frameworks have managed to succeed scientifically, but have largely failed to affect large-scale political and environmental change.

I begin with scientists because global warming has always been first and foremost a problem of science. Since climate change first appeared on the margins of the government’s science policy radar in the 1950s, scientists have served as gatekeepers of information on the Earth’s climate system. Tied to the complex processes of the global atmosphere, the enormous geographical scale and protracted chronology of climate change transcend the limitations of normal human experience.8 Only scientists have the expertise, technologies, and language to understand and communicate the phenomena of

8 See Dipesh Chakrabarty, “The Climate of History: Four Theses,” Critical Inquiry 35 (Winter 2009), 206- 7.

5 this global space. Translating a language of mathematics, data, and models into the everyday narrative of human existence provides lay people with a simulacrum of scientists’ global understanding—an understanding that is itself necessarily based as much on simulation as on direct experience. It is hardly a leap of faith to say that the global atmosphere and its processes exist—they are physical phenomena that occur in real physical spaces with area and volume—but to the extent that environmentalists, politicians, and the public understand global warming and the global atmosphere in their geographical totalities, they understand these things through the medium of science.

Scientists have served as more than simply the gatekeepers of information about the global atmosphere, however; their exclusive access to this global environmental space has all but forced them into roles as advocates. Their advocacy has taken many forms, and it has changed over time.

Initially, few scientists studying the problem understood climate change as a pressing threat to the global environment. Rather, they saw it as a unique opportunity to study and understand what Scripps Institute of Oceanography director characterized as a “large scale geophysical experiment” that might “yield far-reaching insight into the processes determining weather and climate.”9 In the 1950s and ‘60s, their advocacy revolved around promoting the interests of atmospheric science itself.

As I argue throughout this dissertation, climate scientists’ efforts to promote

research and policy on CO2-induced climate change and other atmospheric phenomena arose within dynamic social and political contexts that shaped both the science and the

9 Roger Revelle and Hans E. Seuss, “ Exchange Between Atmosphere and Ocean and the

Question of an Increase in Atmospheric CO2 During the Past Decades,” Tellus 9 (1957): 19.

6 politics of the issue. Common lamentations over the “politicization” of climate science are thus misguided; climate science has always been political.

Nothing shaped the relationship between the science and politics of climate change in the ‘50s and ‘60s more than the Cold War. On one hand, Revelle and his colleagues capitalized on the U.S. government’s Cold War interest in geophysical science in general—and on weather prediction and weather control in particular—to promote a broad research agenda in atmospheric science that included important basic research on

CO2 and climate. They used the potential consequences of both intentional and unintentional climatic change to lobby for greater research funding, first to support individual projects in atmospheric science and later to build the field’s major government-funded institutions, most notably the National Center for Atmospheric

Research (NCAR).

On the other hand, however, scientists’ particular research interests and the institutions they created to pursue those interests also reflected deep anxieties about the

Cold War research system that made their science possible. Scientists emphasized the global nature of their subject matter and stressed the need for international scientific cooperation—an objective they overtly associated with broader international political cooperation in the face of East-West Cold War tension. Meanwhile, at home they designed their institutions—and particularly NCAR—to subvert the hierarchy, secrecy, and centralization that according to Alvin Weinberg’s formulation of “big science” characterized other large, government-funded scientific institutions of the Cold War.10

10 Alvin Weinberg, “Impact of Large-Scale Science on the United States,” Science, vol. 134, no. 3473 (July, 1961): 161-64. See also Weinberg, Reflections on Big Science (Cambridge: MIT Press, 1967); Bruce Hevly and Peter Gallison (eds), Big Science: The Growth of Large-Scale Research (Stanford, CA: Stanford University Press, 1992).

7 Finally, many scientists coupled their commitment to international cooperation and their anxieties about the Cold War research system with a broader concern about the unintended human and environmental consequences of Cold War science itself. It was a concern that atmospheric scientists shared with American environmentalists, and one that allowed the two groups to form a loose alliance in the 1960s based on a mutual interest in

particular types of atmospheric “pollution,” including CO2.

Scientists’ concern over the environmental impacts of climate change grew in tandem with a resurgent American environmental movement in the 1960s and ‘70s, and the complex relationships between these two groups further helped to shape the nature of scientists’ advocacy on global warming. Many atmospheric scientists sympathized with both the precautionary ethos and the specific conservation goals of America’s environmental organizations, and some prominent scientists participated actively in these organizations both as members and as institutional leaders. Increased federal and international interest in environmental affairs during these decades also prompted

atmospheric scientists to frame their research on CO2 and climate change in environmental terms in order to take advantage of new sources of research funding. Both out of a genuine commitment to environmental protection and out of professional interest, in the 1970s scientists helped to transform atmospheric change from a scientific curiosity into a global environmental problem.

Climate change was an atypical environmental problem, however. The same scale and complexity that made climate change the nearly exclusive purview of atmospheric scientists and forced them into roles as advocates also distinguished problems of the global atmosphere from other important environmental issues of the

8 period. The hallmark battles of American environmentalism have historically revolved around specific and discrete geographical spaces. Land use changes, dams, water and common air pollution, pesticides, toxic wastes, and timber and mineral extraction typically affect human and environmental systems on local or regional scales, and environmentalists have capitalized on the local nature of environmental threats to mobilize local, regional, and national political constituencies in grassroots campaigns to protect these geographical spaces. Even the so-called “global environment,” defined by the crisis laid out at the 1972 United Nations Conference on the Human Environment in

Stockholm Sweden, in fact represented an amalgamation of smaller-scale threats to specific environments—threats endemic across the globe but not necessarily global in their action. Problems of the global atmosphere, by contrast—and in particular the threat

of CO2-induced climate change—existed outside of the local, regional, and national geographical frameworks that make other environmental problems governable.

Atmospheric gases like CO2 and ozone circulate almost limitlessly through a boundless

global space. The sources of atmospheric CO2 are diffuse; the gas is a byproduct of fossil fuel energy use, an activity nearly as essential to modern human existence as eating or

drinking. The climatic impacts of CO2, moreover, are divorced from these sources in time and space. Through conferences and reports like the 1970 Study of Critical

Environmental Problems and the 1971 Study of Man’s Impact on Climate, scientists framed atmospheric and climatic change in terms of a larger global environmental crisis.11 But again, the issue only existed in the public consciousness through the

11 These are known commonly as the “SCEP” and “SMIC” reports. Study of Man’s Impact on Climate, Inadvertent Climate Modification: Report of the Study of Man’s Impact on Climate (Cambridge: MIT Press, 1971); Study of Critical Environmental Problems, Man’s Impact on the Global Environment:

9 popularization of atmospheric science. Borderless, complex, and poorly understood, the global atmosphere on its own had no natural political constituency.

If the atmosphere was an atypical environmental space, so too were atmospheric scientists atypical environmental advocates in the 1960s and ‘70s. Environmentalists’ direct political advocacy stood in tension with scientists’ professional commitments to political neutrality and the community-defined standards of “good science.”

Atmospheric scientists recognized that their professional credibility and public authority rested on a loosely articulated set of rules about objectivity, accuracy, and method.

Moreover, the specific issue that introduced the health of the global atmosphere as a meaningful environmental issue and established atmospheric scientists as important players in environmental politics—the debate over the American Supersonic Transport program—involved a broader attack on the kind of Cold War technologies that allowed climate scientists to do their jobs. As the problem of climate change began to look more and more real, these scientists struggled to balance their growing concerns over the

human and environmental consequences of increased CO2 against the professional risks of taking a political stand on the issue. Real scientific disagreement about the severity and even the direction—colder or warmer—of climate change, alongside institutional rivalries, methodological differences, and conflicting personalities only complicated this struggle. Because only scientists had the tools and technologies to address climate change at its truly global scale, in the 1970s these internal politics of climate science defined the broader politics of climate change as an environmental issue.

Assessment and Recommendation for Action: Report on the Study of Critical Environmental Problems (Cambridge: MIT Press, 1970).

10 When climate scientists did engage directly in environmental politics, their particular forms of advocacy on climate change continued to reflect their values and interests as scientists more than they did the middle-class quality-of-life concerns that

Sam Hays describes in Beauty, Health, and Permanence as the philosophical center of the mainstream American environmental movement.12 Environmental organizations relied on broad support from their grassroots constituents as they sought to affect change through political lobbying, litigation, and legislation on specific, typically local

“amenity” issues. Scientists, by contrast, framed climate change in terms of development and natural resources, and they sought to influence elites in government and at international scientific organizations more than they worked to mobilize the American public. At home, they focused on the potential consequences of climate change on water resources, agricultural production, and national defense. Amidst a series of international climatic anomalies in the 1970s, climate scientists also hoped to investigate the relationship between climate change and essential resources like food, water, and other natural resources abroad. Above all, throughout the political history of global warming, climate scientists continued to promote better science as a route to better policy. Good science, they believed, in the hands of rational political actors, would pave the way for reasonable and progressive solutions to an increasingly clear global environmental problem.

12 After 25 years, Samuel P. Hays’ Beauty, Health, and Permanence: Environmental Politics in the United States, 1955-1985 (New York: Cambridge University Press, 1987) still provides the most comprehensive analysis of post-war American environmentalism. See also Robert Gottlieb, Forcing the Spring: The Transformation of the American Environmental Movement (Washington, D.C.: Island, Press, 1993); Hal K. Rothman, The Greening of a Nation?: Environmentalism in the United States Since 1945 (Fort Worth, TX: Harcourt Brace & Company, 1998); Kirkpatrick Sale, The Green Revolution: The American Environmental Movement, 1962-1992 (New York: Hill & Wang, 1993); Philip Shabecoff, A Fierce Green Fire: The American Environmental Movement (revised edition) (Washington, D.C.: Island Press, 2003).

11 And initially, the strategy seemed to work. By the end of the 1970s, scientists had gained remarkable access to government resources and officials in government agencies sympathetic to their calls for more research on climate change and open to an eventual plan to incorporate global warming and its consequences into the nation’s long-term energy strategies. Indeed, despite divisions within the scientific community and a persistent disconnect between scientists and American environmentalists, a snapshot of climate science and climate change politics in the late 1970s could not help but inspire a certain optimism about society’s capacity and willingness to tackle the problem of global warming in the 1980s.

But scientists’ commitment to working within the federal government to influence policy through what one American Association for the Advancement of Science leader called the “forcing function of knowledge” left the climate science community profoundly vulnerable to political change.13 Not only did scientists turn to government agencies like the Department of Energy and the Council on Environmental Quality to incorporate their research into national level energy planning and environmental policy, they also relied on these and other extensions of the executive branch for the material and financial resources to conduct their science in the first place. With the exception of a few controversial popularizations of climate change aimed at the general public, even the most politically active scientists shied away from the kind of value-based arguments that environmentalists used to build a popular political constituency. With little public backing and no recourse to the type of legal and political strategies that environmentalists used as a check on objectionable government policy, in the early 1980s scientists lacked

13 Draft notes of AAAS Advisory Group on Climate meeting, May 26, 1978, AAAS Working Group on Climate Meeting (Transcript), AAAS Climate Program Records, AAAS Archives, Washington, D.C.

12 the means and the clout to defend themselves—and their science—against a new

Presidential administration hostile to both environmental regulation and environmental science.

The “Reagan Reaction” exposed the liabilities of scientists’ preferred forms of scientific and political advocacy. As many scientists and environmentalists bitterly recall, the new President actively and systematically weakened all sorts of federal environmental protection programs from regulations on mining and logging to research on renewable energy. The Administration replaced capable administrators at the EPA,

CEQ, DoE, and the Office of Science and Technology Policy with loyal political acolytes who endeavored to undermine the missions of the very government bodies they led. The

White House slashed government funding for social and environmental science research,

including the Department of Energy CO2 research program meant to serve as the primary interface between climate science and federal energy policy under President Carter. As

Stephen Schneider remembers in Science as a Contact Sport, many of these administrators not only cut climate scientists’ funding; they also attacked these scientists’ credibility and challenged the validity of their conclusions on global warming.14 In denying climate scientists both resources and access to government officials, the

Administration rendered scientists’ established forms of scientific and political advocacy effectively moot.

While the Reagan Reaction fundamentally changed the nature of global warming politics in the 20th century, however, that change came as much from scientists’ and environmentalists’ responses to the Administration as it did from the Administration

14 Schneider, Science as a Contact Sport, 84-95.

13 itself. Facing a common foe, climate scientists began to work with American environmental organizations and Congressional Democrats like Congressman George

Harrison and Senators Tim Wirth and to undercut the President’s science and environmental policies. Politically active climate scientists like Schneider and Roger

Revelle also raised the stakes of the discussion on global warming by reframing their atmospheric research in terms of the President’s controversial and environmentally destructive policies on energy and nuclear defense. Their activism helped to establish global warming as a mainstream political issue associated with a vocal minority on the political left. Working together, scientists, environmentalists, and Democrats thus reshaped the politics of global warming as a politics of dissent.

This new coalition faced an old set of problems in a new political landscape.

Environmentalists had responded slowly and tepidly to the issue of climate change in the

1970s in part because they feared that at its truly global scale, the problem would explode the local, regional, national, and even international political mechanisms that environmentalists had established to affect change on other pressing environmental issues in the 20th century. And it did. Few legal or political tools existed at any level of government to regulate greenhouse gas emissions in the 1980s, and leaders in climate science and at environmental organizations found themselves scrambling to create new international legal and political strategies that could cope with large-scale problems of the global atmosphere. Along with international leaders at the United Nations, advocates of action on global warming turned to their own successful efforts to deal with two related atmospheric problems in the early 1980s—acid rain and ozone depletion—for guidance.

Their collective successes in establishing international legal conventions on acid rain and

14 ozone—and particularly the Convention on the Protection of the Ozone Layer and the Montreal Protocol on Ozone Depleting Substances—not only provided these groups with the confidence that such frameworks could be built to handle greenhouse gases, but also with a rough template for how to build them.

The stakes of an international convention on global warming were much higher than those of a convention on ozone, however. Protecting the ozone involved eliminating a few relatively easily replaceable chemicals from non-essential consumer products like

aerosol cans and air conditioners. Regulating CO2 and climate change required fundamental changes to the energy and land use patterns at the heart of the global economy.

As they had with ozone, advocates of a convention on global warming focused first and foremost on establishing a strong international scientific consensus to support an

international legal framework. But because of the relationship between CO2 and energy, their efforts to create a consensus on science for the purpose of policy met stiff opposition from powerful representatives of national governments and energy-dependent industries. Political leaders ignored or rejected traditional forms of consensus that gave them no say in the contested scientific facts at hand. In response, global warming advocates established a new, intergovernmental form of politically negotiated scientific consensus that included not only scientists, but also representatives from the national governments, industries, and environmental NGOs whose participation would be necessary to turn science into meaningful international policy. In 1988, this process was formalized under the Intergovernmental Panel on Climate Change (IPCC).

15 In its original form, the IPCC was not a mechanism for mitigating climate change or even for making climate policy; rather, the IPCC was mechanisms for making a certain type of knowledge. More specifically, scientists and United Nations officials intended the IPCC as a way to provide scientific guidance for a new United Nations Framework

Convention on Climate Change (UNFCCC), an international treaty introduced at the

1992 United Nations Conference on Environment and Development in Rio de Janeiro,

Brazil, or the “Rio Earth Summit.” The distinction between the science of the IPCC and the politics of the UNFCCC was hardly a clear one; on the contrary, scientists actually designed the IPCC as a self-consciously political process of presenting and verifying scientific information. It was not a politics that scientists could easily control, however.

Because the design, function, and ratification of the UNFCCC relied on the substance and certainty of the scientific conclusions of the IPCC, the politics of consensus-making on global warming quickly became the politics of global warming itself. Almost as soon as it was formed, the IPCC thus served not only as the primary forum for making knowledge about climate change, but also as the primary locus of climate change politics more broadly, masked in the language of science.

Making Global Warming Green concludes with the formation of the IPCC and the

UNFCCC, although clearly the political history of global warming does not stop there. In part, closing this history in 1992 is a matter of practicality. As global warming has continued to grow as a political, economic, social, and scientific issue in the past twenty years, the amount of information available to the historian has expanded almost exponentially. At the same time, certain official documents—Congressional and

Presidential records among them, but also official documents of environmental

16 organizations like the Sierra Club and the World Resources Institute and politically conservative institutions like the Marshall Institute and the Global Climate

Coalition—are not yet available. The size of the historical record continues to grow; the quality of its organization and the reliability of its contents remains relatively poor. So while a comprehensive political history of global warming up to the present may yet be possible, it is simply beyond the scope of a dissertation.

In part, however, my periodization is also a matter of argument. This is a descriptive work, focused on the development of the structures of global warming advocacy in the 20th century. I contend that this dissertation is valuable to understanding the contemporary problem of global warming because the forms of political advocacy developed by scientists and environmentalists between the 1950s and the early 1990s continue to dominate global warming politics today. The scale and complexity of the problem continue to divorce global warming from everyday human experience and explode the normal mechanisms of national and international policy. Scientists and their forms of knowledge continue to dominate global warming discourse, and science itself continues to serve as a medium for its politics.

But I also contend that the persistent intellectual, institutional, and political structures if global warming politics are inseparable from the historical contexts in which they were developed. Moreover, it is easier and perhaps more productive to try and understand the political history of global warming when we distance ourselves from the actors, institutions, and events of that history in time. Because of its chronological proximity, historians in general tend to treat the second half of the 20th century as a sort of prelude to the present. But many of the related political realities that shaped the science

17 and politics of climate change—the aerospace technology boom of the 1960s, the birth of modern American environmentalism, and perhaps more than anything, the ever-changing contours of the Cold War—have either changed so radically as to be unrecognizable in the 21st century, or have disappeared altogether. From the beginning, the influences of these 20th century political realities—and especially of the Cold War—have had consequences for the both the science and politics of global warming as the issue has evolved over time. As L.P. Hartley famously wrote, “the past is a foreign country; they do things differently there.”15 So too is the 20th century a foreign country. It is where the politics of global warming were born.

15 L.P. Hartley, The Go Between, (Middlesex: Penguin Books, 1961).

18 Chapter 1 The Cold War Roots of Global Warming

In 1938, a British steam engineer and amateur named Guy Stewart

Callendar read a paper before the Royal Meteorological Society, his first before that esteemed body. Confident and athletic despite his almost comically long legs, the 40 year-old Callendar argued that human consumption of fossil fuels was, via the

accumulation of CO2 in the atmosphere, slowly raising the mean temperature of the

Earth. Fossil fuel combustion, he showed, had added about 150 billion tons of CO2 to the atmosphere, about 2/3 of which remained aloft. Using the known radiation absorption

coefficients for water and CO2, he estimated that this increase in atmospheric CO2 should cause the world to warm at a rate of about .003ºC per year—a figure not far from the

.005ºC rate of warming measured by in the half century before

Callendar’s paper.16 “Few of those familiar with the natural heat exchanges of the atmosphere,” he wrote, “would be prepared to admit that the activities of man could have any influence upon phenomena of so vast a scale.”17 Nevertheless, Callendar contended,

“such influence is not only possible, but is actually occurring at the present time.”18 The

Earth, he declared, was warming, and humans were responsible.

Callendar’s contention about anthropogenic warming now seems both prescient and remarkable, but his work initially garnered little attention outside of a small community of meteorologists, and even they expressed skepticism about his conclusions.

Callendar’s thesis—a revised version of an existing and unpopular CO2 theory of climate—rested on three primary contentions: 1) that atmospheric CO2 was rising; 2) that

16 G.S. Callendar, “The Artificial Production of Carbon Dioxide and its Influence on Temperature,” Quarterly Journal of the American Meteorological Society, 64 (1938):223. 17 Callendar, “Artificial Production,” 223. 18 Ibid.

19 increased atmospheric CO2 would lead to warming; and 3) that the Earth was in fact getting measurably warmer.19 Eventually, scientists would vindicate Callendar on each of

these points, and by as early as 1965, research on CO2 and climate stood high on the list of priorities for the growing field of atmospheric science.20 At the outset of the Second

19 The CO2 Theory of Climate has a long history before G.S. Callendar, beginning as early as 1822 with the work of Jean Baptiste Joseph Fourier, but also including work by John Tyndall in the mid-19th century, Swedish Nobel Laureate Chemist and American Geologist T.C. Chamberlain at the turn of the century, as well as many others. The story of the CO2 theory of climate in fact presents an interesting microcosm of the great transformation of the geophysical sciences that occurred in the 19th century. See James Roger Fleming, Historical Perspectives on Climate Change (Oxford: Oxford University Press, 1998) and The Callendar Effect: The Life and Times of Guy Stewart Callendar (Boston: American Meteorological Society, 2007); Joseph Fourier “Remarques générales sur les témperatures du globe terrestre et des espaces planétaires,” Annales de Chimie et de Physique (Paris), 2nd Ser., 27 (1824): 136-67, published in English in 1837 as “General Remarks on the Temperatures of the Globe and the Planetary Spaces,” trans by Ebeneser Burgess, American Journal of Science, v. 32 (1837): 1-20; I. Grattan- Guiness, Joseph Fourier, 1768-1830: A Survey of His Life and Work (Cambridge: MIT Press, 1972); John Herivel, Joseph Fourier: The Man and the Physicist (Oxford: Clarendon Press, 1975); John Tyndall, “On Radiation Through the Earth’s Atmosphere,” Philosophical Magazine, ser. 4, 125 (1862): 200-206; John Tyndall, “On the Absorption and Radiation of Heat by Gases and Vapours, and on the Physical Connection of Radiation, Absorption, and Conduction,” Philosophical Magazine, ser. 4, 22 (1861): 169-94, 273-85; Svante Arrhenius, “On the Influence of Carbonic Acid in the Air Upon the Temperature of the Ground,” The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science, ser. 5 (April 1896): 237-276; Arrhenius, Worlds in the Making: The Evolution of the Universe, trans. by H. Borns (New York: Harper and Brothers, 1908); Naomi Oreskes and Ronald E. Doel, “The Physics and Chemistry of the Earth,” in The Cambridge History of Science Volume V: Modern Physical and Mathematical Sciences, ed. Mary Jo Nye (Cambridge: Cambridge University Press, 2002), 538-560. For more on theories of climate change in the 19th and early 20th centuries, see John Imbrie and Katherine Palmer Imbrie, Ice Ages: Solving the Mystery (Cambridge: Harvard University Press, 1979); A. Berger, “Milankovitch Theory and Climate,” Review of Geophysics 26 (1988): 624-57; C.E.P. Brooks, Climate Through the Ages: A Study of the Climatic Factors and Their Variations, 2nd Revised Edition (New York: Dover Publications, 1970). 20 Atmospheric CO2 was in fact, by Callendar’s measurement, on the rise, but geophysical scientists differed greatly on the rate at which the excess of CO2 might be absorbed by the oceans or other sinks in the carbon cycle, complicating any prediction of future atmospheric CO2. Estimates of CO2 uptake varied by a factor of 100. The radiative properties of CO2 itself, meanwhile, also had yet to be fully hashed out. Some scientists argued that the infrared absorption bands for CO2, H20, and other gases were sufficiently narrow to absorb all of the available infrared energy at those wavelengths, meaning that beyond a certain point, additional CO2 would have no effect on infrared radiation. The temperature trend, too, was subject for debate. In the early 1940s, some climatologists argued that the gradual increase in temperature represented no more than a “casual” natural variation in climate. By the early 1950s, the persistent warming trend began to filter into the public consciousness, but just as it became fodder for media speculation and pithy cartoons, the trend peaked, leveled off, and began to descend. Callendar’s predictions fell flat. Perhaps more importantly than the failures of these predictions, however, was the accuracy of Callendar’s prediction that few scientists would believe that humans had the capacity to alter the climate at such a large. As U.S. Weather Bureau climatologist Helmet Landsberg wrote authoritatively in Scientific Monthly in 1946, “the outdoor climate cannot be changed, except on the smallest scale.” Helmut E. Landsberg, “Climate as a Natural Resource,” The Scientific Monthly, Vol. 63, No. 4 (Oct. 1946): 293. For the best source on Callendar, see Fleming, The Callendar Effect. See also Fleming, Historical Perspectives.

20 World War, however, Callendar’s CO2 theory of climate remained, like Callendar himself, relatively obscure.

Between 1945 and 1965, the Cold War provided a way for scientists to bring

21 research on CO2 and climate into the scientific and political mainstream. The detonation of the first plutonium bomb in the desert near Alamogordo, New Mexico, in

1945 presaged an important philosophical change in the way scientists and the public understood humans’ relationship with the natural world, and this transformation was essential in the making of modern climate science. The atomic bomb established humans as agents of large-scale geophysical change. It opened the door for the possibility that human actions might already be altering geophysical processes unintentionally, just as

Callendar had argued. Perhaps more importantly, the advent of the bomb also fostered the development of an expansive Cold War research system that provided the financial and technological resources necessary for studying the Earth’s large-scale geophysical

systems. Though the CO2 theory of climate predates the Cold War by over a century, it is

21 Historians of global warming have typically treated the Cold War as a sort of backdrop for the advance of scientific knowledge—one “lens” of many through which to understand the history of climate science, as Spencer Weart notes in The Discovery of Global Warming (Cambridge: Harvard University Press, 2003) and elsewhere. And indeed, it is important to remember that the Cold War provided the milieu in which the science and politics of global warming developed through its first thirty-five years. The relationship between global warming and the Cold War is perhaps more complex than this, however; it involves specific topical, personal, and institutional connections between the study of CO2 and the prosecution of America’s half-century scientific and military competition with the . Few scholars have tackled these relationships in depth, but there are exceptions. Weart’s analysis of the related popular imagery on the two problems in the Bulletin of the Atomic Scientists provides an interesting starting point, although here as in Discovery, he does more to point out that popular and political concerns about warming and nuclear energy were related than he explains the specific nature of that relationship. David Hart and David Victor do a better job by following the money in their presentation of atmospheric scientists’ efforts to promote CO2 research as a model of Cold War scientific entrepreneurship. David Hart and David Victor, “Scientific Elites and the Making of U.S. Policy for Climate Change Research, 1957-1973,” Social Studies of Science, Vol. 23, No. 4 (Nov., 1993), 643-680. Much work remains to be done on the Cold War roots of global warming, however (hence this chapter), and I will revisit the relationships between the Cold War, climate science, and American environmentalism throughout the dissertation, especially in chapters 5 and 6.

21 the incorporation of atmospheric science into this Cold War research agenda in the 1950s and ‘60s that marks the beginning of the modern history of global warming. 22

A handful of powerful individuals drove the growth of Cold War atmospheric

research in the 1950s, and these entrepreneurial scientists integrated studies of CO2 and climate into the larger Cold War research agenda in two main ways. First, individual

scientists interested in CO2 and related issues capitalized on existing government research efforts to gain funding and support for specific projects that involved measuring and

monitoring atmospheric constituents like CO2. Research into nuclear testing, geophysical

modification, and fallout led to advances in scientists’ knowledge of CO2, atmospheric circulation, and climate; scientists in turn framed these climate-related subjects as important elements in the study of militarily significant issues like the distribution of radioactivity, weather and climate control, and atmospheric monitoring. The potential military applications of weather prediction and weather control complemented persistent domestic demands for better local meteorological services, and as the field grew in the

1950s and ‘60s, atmospheric scientists found their work increasingly relevant to both domestic and foreign policy needs.

22 In the past two decades, historians of global warming have joined scientists and popular authors in creating a sort of standard narrative for the history of global warming that reaches back into the 19th Century. If there is a single, standard text on this history, it is Spencer Weart’s The Discovery of Global Warming, which provides a basic history of the scientific developments and discoveries that led to the modern science of global warming. For a very readable history of climate science that traces the field’s meteorological roots, see William K. Stevens, The Change in the Weather: People, Weather, and the Science of Climate (New York: Dell Publishing, 1999); for another popular history worth reading, see Gale E. Christianson, Greenhouse: the 200-year Story of Global Warming (New York: Penguin Books, 1999). James Roger Fleming offers perhaps the most critical and nuanced historical work on the science of climate, both in myriad articles on weather modification and in Historical Perspectives. Fleming is careful to note how the ideas about the concept of climate—its scale, its impact, and its permanence—have changed over time, although he, like other historians of science, has a tendency to focus on the details of “first references” and “discoveries,” sometimes to the detriment of his discussion of these events’ historical significance.

22 Second, leaders of and climatology sought to capitalize on the government’s interest in geophysical research to promote their disciplines through government-sponsored institutions, and in particular through the National Center for

Atmospheric Research (NCAR) in Boulder, Colorado. Founded nearly a decade after the mainstay institutions of post-war government research—the National Science

Foundation, the Atomic Energy Commission, the National Research Council, and a variety of government-sponsored defense laboratories—NCAR relied on the Cold War science establishment for financial support.

At the same time, however, the institution’s incorporation of domestic scientific concerns, its relaxation of hierarchical management structures, and its emphasis on the needs of the academic community also reflected scientist’s ambivalence about the structures of that establishment.23 Atmospheric scientists’ uneasiness about Cold War

23 Historians have spent a great deal of ink discussing the structures of Cold War science, and though the focus has primarily been on physics, this historiography provides an interested lens through which to view the development of atmospheric science in the 1950s and ‘60s. Much of the discussion has revolved around scientists’ autonomy in government-sponsored research. In 1987, Paul Foreman argued that physicists maintained only the “illusion of autonomy” while pursuing basic research sponsored by government sources, and that while they may have benefited from government resources, these scientists were used by American government and society in the prosecution of the Cold War far more than they “used” the structures of state-sponsored science. Daniel Kevles, however, reached the opposite conclusion in his 1990 essay, “Cold War and Hot Physics: Science, Security, and the American State.” As Joseph Manzione put it in a recent review essay for the journal Diplomatic History, Kevles “found that the integration of physicists into administration and government during the fifties brought welcomed diversification and resources to the discipline and a measure of power to the scientists to define their own research agenda and to influence policy.” As Manzione notes, most recent scholarship has upheld Kevles point of view, but this is hardly the whole story. Manzione points to another tension inherent to the government sponsored Cold War research system: the tension between the internationalist values of American scientists and the national and often military framework in which they operated. During the interwar years, many scientists espoused a form of “scientific internationalism” that put the interests of science above any sort of non-scientific political allegiance, and Manzione characterizes the eagerness with which American scientists took on war work as a violation of this basic principle. Gregg Herkin takes a second look at Manzione’s characterization of Cold War science as a sort of Faustian bargain in his own Diplomatic History article of 2000. For Herken, the analogy does not hold, for there is certainly a difference between selling your soul to the devil and signing on for the mutual interest of science and the democracy that you as a citizen have a stake in upholding. Finally, Aaron L. Friedberg’s 1996 review essay, “Science, the Cold War, and the American State” tackles the Foreman-Kevles dichotomy from a more productive angle—and one that is relevant to the history of the atmospheric sciences, not only in the 1950s, but through the end of the Cold War. Friedberg generally agrees with Kevles’ analysis of a Cold

23 research betrayed a mixture of optimism and fear about scientific and technological development more generally—a mixture also at the heart of America’s nascent environmental movement in the 1960s.

The relationship between atmospheric scientists and environmentalists was loose at best; where it existed, it grew primarily from a mutual concern about various forms of

air pollution. As their early attempts to frame atmospheric CO2 as a unique type of pollutant reveal, however, in the 1960s some atmospheric scientists began to espouse an ethos of precaution that they shared with biological scientists and ecologists more closely associated with the American environmental movement. Scientists and politicians continued to promote the ability of modern science to transcend the limitations of nature

and solve human problems, but by 1965, CO2-induced warming brought on by human consumption of fossil fuels had been identified as one of the disconcerting man-made problems that science might have to set out to solve.

War research structure that that was by and large “diffuse and pluralistic.” Nevertheless, he argues, there was a constant tension between the drive for centralized, militarily-applicable research in the service of Cold War competition and a more popular, democratic ideal of scientific administration that fit with American democratic more generally. He points out scientists’ self-serving ideological commitment to the “free market” in the government research contract—a category of award that was in fact only rarely doled out in an arena of open, unfettered competition. Ultimately, this same set of concerns—concerns over scientific autonomy, over scientific internationalism, and over scientific centralization—dominated the development of atmospheric science during the Cold War. More to the point, these were tensions endemic to the Cold War research system itself—in fact, they were some of the structures of Cold War science. Paul Forman, “Behind Quantum Electronics: National Security as a Basis for Physical research in the United States,” Historical Studies of the Physical and Biological Sciences 18, No. 1 (1987): 149-229; Daniel J. Kevles, “Cold War and Hot Physics: Science, Security and the American State,” Historical Studies of the Physical and Biological Sciences, Vol. 20, No. 2 (1990): 239-64; Aaron L. Friedberg, “Science, the Cold War, and the American State,” Diplomatic History, Vol 20 (Winter, 1996): 107-118; Joseph Manzione, “Amusing and Amazing and Practical and Military: The Legacy of Scientific Internationalism in American Foreign Policy, 1945-1963,” Diplomatic History, Vol. 24, No. 1 (Winter, 2000): 21-55; Gregg Herken, “In the Service of the State: Science and the Cold War,” Diplomatic History, Vol. 24, No. 1 (Winter, 2000): 107-115. See also Daniel Kevles, The Physicists: The History of a Scientific Community in Modern America (Cambridge: Harvard University Press, 1995).

24 Roger Revelle, CO2 and the IGY

Initially, the links between research on CO2 and climate and the larger U.S. Cold

War research system revolved around a few individual scientific leaders whose interest in

CO2 overlapped with related research in other geophysical sciences like oceanography and geochemistry. These wide-ranging entrepreneurial scientists capitalized on the discoveries, technologies, and most of all the resources made available by the government’s heightened interest in studying geophysical processes, especially as they applied to questions of the distribution and circulation of radioactive isotopes and nuclear fallout. Collectively, these scientists helped to define the research priorities for

atmospheric science in the decades to come. One of these priorities was research on CO2.

No individual did more to put atmospheric CO2 on the Cold War research agenda in the 1950s and early ‘60s than the self-appointed “‘granddaddy’ of the theory of global warming,” Roger Revelle.24 Born in Seattle, Washington in 1909, Revelle spent most of his life in Southern , where he attended Pomona College, married into the wealthy and well-connected Scripps family, and eventually earned a graduate degree from, worked as a professor and research scientist at, and became the director of the

Scripps Institute of Oceanography. Tall, fit, adventurous, and by all accounts wonderfully charismatic, he personified the entrepreneurship of the scientific leaders of the post-war period. A navy officer and the director of one of the nation’s most important centers of oceanography, he was instrumental in the development of the post-

24 The appellation “grandfather,” “godfather,” or “father” of global warming shows up in all sorts of places, from interviews with Revelle himself to accounts of his life. See Walter Sullivan, “Roger Revelle, 82, Early Theorist in Global Warming and Geology,” New York Times, July 17, 1991; Judith Morgan and Neil Morgan, Roger: A Biography of Roger Revelle (San Diego: University of California, San Diego—Scripps Institution of Oceanography, 1996); Deborah Day, Roger Randall Dougan Revelle Biography (UC San Diego: Scripps Institution of Oceanography Archives, 2008), http://escholarship.org/uc/item/78d9v14v; Fleming, Historical Perspectives,122.

25 war Office of Naval Research (ONR), an outgrowth of the wartime National Research

Defense Council and the Navy Office of Research and Inventions and an organization that was responsible for funding and oversight for the vast majority of government- sponsored geophysical science between the end of WWII and the establishment of the

National Science Foundation (NSF) in 1950.25 Revelle was good at finding money for both his own and his colleagues’ research. His work on climate and climatic change in

the 1950s helped not only to rekindle mainstream interest Callendar’s CO2 theory of climate; it also generated financial and material support for some of the most fundamental research on climate in the ensuing decades.

Revelle’s place in the history of climate change has been the subject of some debate, however. In the popular press and among many scientists, he serves as a benevolent progenitor of modern climate science. In An Inconvenient Truth, for example, he is cast not only as a sort of climatic Paul Revere, but also as a mentor to then student and future Senator, Vice President, and Nobel Laureate Al Gore.26 But some historians of science feel that his role as been overblown. In particular, James Roger Fleming

questions both the novelty and intent of Revelle’s early pronouncements on CO2, citing

“Revelle’s need to place himself at the center of the carbon dioxide theory of climate as a way of maintaining his larger-than-life legend.”27 Fleming is perhaps the most vocal historical champion of the too often forgotten Guy Stewart Callendar, and he points out

Revelle’s cavalier appropriation of parts of Callendar’s theory as a way to at once elevate

25 Deborah Day, Roger Randall Dougan Revelle Biography (not paginated), Thomas F Malone, Edward D. Goldberg, and Walter H. Monk, Roger Randall Dougan Revelle, March 7, 1909-June 15, 1991, National Academy of Sciences Biographical Memoirs, www.nasonline.org; Roger R. Revelle, "Preparation for a Scientific Career," an oral history by Sarah Sharp, 1984, Regional Oral History Office, The Bancroft Library, University of California, Berkeley, 1988. S10 Reference Series No. 88, November, 1988. 26 An Inconvenient Truth, directed by Davis Guggenheim (Hollywood, CA: Paramount Pictures, 2006). 27 Fleming, Historical Perspectives, 128.

26 Callendar and knock Revelle off of what Fleming sees as his largely self-constructed pedestal.28 The argument has some merit; in the grand scheme of things, Revelle

“discovered” relatively little in his work on CO2 and climate. He relied heavily on the work of others, including Callendar.29 Nevertheless, as a scientific administrator with ties to scientists and science policymakers in private companies, universities, government agencies, and international organizations, Revelle’s early and substantial support for

research into CO2 and climate put these issues on the Cold War science map.

Revelle’s most important scientific contribution to the study of climate involved a reevaluation of the role of the oceans in the global carbon cycle, a study supported by funds from ONR and the Atomic Energy Commission (AEC).30 In the early 1950s, nuclear weapons tests dumped large, known amounts of radiation into the atmosphere at specified times and places, creating a “ready made experiment for tracing the circulation of carbon” through the atmosphere.31 The AEC and ONR subsequently began to take a keen interest in radioactive carbon…and in where it went. Amidst nascent concerns about nuclear fallout from bomb tests at Bikini Atoll (where then Commander Revelle had headed a team studying a coral lagoon in 1946) and elsewhere, the government began to ask oceanographers like Revelle just how fast the oceans could swallow up the radioactive carbon that the U.S. military was putting into the atmosphere.32

28 Fleming, Historical Perspectives, 122-124; Fleming, The Callendar Effect, 79-80. 29 Revelle’s major scientific contribution to global warming involved the chemistry and mixture of sea- water, as discussed below. 30 Revelle’s interest in the carbon cycle dates back to his work on carbon deposits in the ocean for his dissertation, "Marine Bottom Samples Collected in the Pacific Ocean by the CARNEGIE on its Seventh Cruise.” Day, Roger Randall Dougan Revelle. 31 Hart and Victor, 648. 32 For more on Revelle’s involvement in Operation Crossroads and other nuclear testing activities, see Roger Randall Dougan Revelle, “Preparations for a Scientific Career,” Interviews by Sarah L. Sharp, 1984, part 1, Scripps Institution of Oceanography Archives, , CA (copyright 1988, Regents of the University of California); Roger Revelle, interview by Earl Droessler, February 1989, American Institute of

27 Working with Hans Suess, an Austrian-born physical chemist, nuclear physicist, and radiocarbon dating expert who he hired on at Scripps with ONR/AEC funding in

1955, Revelle began to use measurements of radiocarbon in air and seawater to

33 investigate the rate of CO2 exchange between the oceans at the atmosphere. At the

United State’s Geological Survey’s laboratory in Washington, D.C., Suess had made preliminary measurements of radioactive carbon isotopes in ancient trees and compared them to measurements of carbon in the atmosphere. He detected from the ratio of carbon-14 to carbon-12 that ancient carbon from the burning of fossil fuels had diluted

the “natural” CO2 of the atmosphere—a radiochemical confirmation of Callendar’s earlier

34 contention that fossil fuel use contributed CO2 to the atmosphere.

At first, Revelle and Suess both assumed—like many other scientists—that the

oceans would absorb this excess CO2, and together they hoped to study and trace

Physics, College Park, MD; Laura Harkewicz, “Oral History of Gustaf Olof Arrhenius,” April 11, 2006, Scripps Institute of Oceanography Archives, http://libraries.ucsd.edu/locations/sio/scripps- archives/resources/collections/oral.html; See also Weart, 28. 33 Suess is an interesting character in the history of science whose life merits further study. The third generation in an Austrian scientific dynasty, Suess received his Ph.D. in 1936, and worked for the German government on “”—that is, water rich in the isotope deuterium—and on various radiation- related problems during the war. He was a not particularly renown member of the cadre of German scientists brought to the U.S. after the war during the competition between the Americans and Soviets for Nazi scientific brain power, and only after working at the Fermi Institute at the did he eventually move to the USGS and on to Scripps. For more, see Heinrich Waenke and James R. Arnold, “Hans E. Suess, 1909-1993,” NAS Biographical Memoirs, http://www.nasonline.org. 34 As plants grow, they “fix” the carbon from atmospheric CO2 as organic material, including a naturally occurring radioisotope called carbon-14. In 1949, Willard Libby of the University of Chicago devised a method for measuring the rate at which carbon-14 decays, giving archeologists and paleontologists a new way of dating very old relics and the dirt they came from. Suess applied the idea to geochemical measurement, and calculated what the overall ratio of carbon-14 to carbon-12—that is, normal carbon—should be, given a set of approximate time scales for major influences on CO2 like weathering and ocean mixing. When the ratio was lower than he had anticipated, he demonstrated that the whole mix had been diluted by carbon from ancient sources, and particularly from fossil fuels, which contain little if any Carbon-14. This dilution is known as the “Suess effect.” For more on Suess and his effect, see Hans E. Suess, “Radiocarbon Concentration in Modern Wood,” Science, Vol. 122 (2 September 1955), 415-17; Keeling, C.D. (1979), “The Suess effect: 13Carbon-14Carbon interrelations.” Environment International 2: 229–300; “Hans Suess Papers: Background,” from the University of California San Diego Library Special Collections website, http://orpheus.ucsd.edu/speccoll/testing/html/mss0199d.html; Fleming, Historical Perspectives, 125; Weart, Discovery, 28-29.

28 radioactive CO2 as it was absorbed by and circulated through the seas. Shortly before they sent their paper on the subject to the journal Tellus in 1957, however, it occurred to

Revelle that they had failed to account for the tendency of sea-water—a complex chemical stew, to paraphrase Spencer Weart—to retain a generally constant acidity

35 through a self-regulating “buffering” mechanism involving CO2. The CO2 that most scientists assumed would be absorbed might just as easily be re-released while still at the ocean surface, resulting, as Callendar had argued, in an overall increase in atmospheric

CO2—an increase with interesting and far-reaching geophysical significance. Revelle’s realization about buffering was a late addition to the Revelle-Suess paper, which in the main actually challenged the extent of the so-called “Callendar Effect.” Nevertheless, the paper—and, more importantly, its authors—made an impact. “Human beings,” the scientists famously wrote,

“are now carrying out a large scale geophysical experiment of a kind that could

not have happened in the past nor be reproduced in the future. Within centuries

we are returning to the atmosphere and oceans the concentrated organic carbon

stored in sedimentary rocks over hundreds of millions of years. This experiment,

if adequately documented, may yield far-reaching insight into the processes

determining weather and climate.”36

35 Weart, Discovery, 28. 36 Roger Revelle and Hans E. Suess, “Carbon Dioxide Exchange between Atmosphere and Ocean and the

Question of an Increase in Atmospheric CO2 during the Past Decades,” Tellus 9 (1957): 19. This , or parts of it, has been quoted as the starting point for scientists’ concern over global warming, though as Fleming discusses in detail, both the specific subject and the tone of the passage echoes earlier statements by other scientists, particularly Guy Stewart Callendar and Gilbert Plass. See Gilbert N. Plass, “The Carbon Dioxide Theory of Climate Change,” Tellus VII (1956), 2: 140-154; See also Fleming, Historical Perspectives, 107-128; Weart, Discovery, 30.

29 For Revelle and Suess, the only immediate danger in this grand experiment was that it might not be adequately monitored and documented. 37 Revelle took steps to ensure that it was. Taking advantage of resources available to him as the president of the

Scientific Committee on Ocean Research (SCOR), a committee of the International

Council of Scientific Unions (ICSU), Revelle designed an atmospheric monitoring program for the 1957-58 International Geophysical Year (IGY).38 In July of 1956,

Scripps hired , a geochemist focused on measuring atmospheric

CO2, and in 1957 Revelle set him up with IGY funds and put him in charge of the new

39 IGY atmospheric monitoring program. Keeling established CO2 monitoring stations at the Mauna Loa Observatory in Hawaii and at a research post in Antarctica in order to

establish a baseline measurement of atmospheric CO2 that could be used to measure

40 future changes. In 1957, the amount of CO2 in the atmosphere across the globe stood at

37 The statement, often attributed to Revelle alone, has also been framed as an early warning on global warming, though it seems to have been little more than a statement of scientific interest at the time. In a beautifully annotated oral history of Gustaf Arrhenius, the son of Svante Arrhenius and a colleague of Revelle at Scripps, the subject remembers that “Roger wasn’t alarmed at all either—he liked great geophysical experiments. He thought that this would be a grand experiment to make, if possible, particularly because of his oceanographic background—to study the effect on the ocean of the increase of carbon dioxide in the atmosphere and the mixing between the ocean reservoirs.” “Oral History of Gustaf Olof Arrhenius,” conducted by Laura Harkowitz, April 11, 2006, La Jolla, CA, Regents of the University of California, 2006; Revelle and Suess, “Carbon Dioxide Exchange,” 19; Roger Revelle, testimony in US House of Representatives, Committee on Appropriations, National Science Foundation - International Geophysical Year (1956), 473. See also Roger Revelle’s testimony in U.S. Congress, Report on International Geophysical Year, House Committee on Appropriations, May 1, 1957, 85th Congress, 1st Session (Washington, D.C.: U.S. Government Printing Office, 1957): 113. 38 Hart and Victor, “Scientific Elites,” 651. 39 Deborah Day, Roger Randall Dougan Revelle Biography (UC San Diego: Scripps Institution of Oceanography Archives, 2008), http://escholarship.org/uc/item/78d9v14v. Money continued to be a struggle after the IGY for Keeling and his Mauna Loa research station, and in 1958, Revelle went so far as to divert funds granted to Scripps by the Atomic Energy Commission in Keeling’s direction. Weart, Discovery, 36. 40 Keeling was picked in part because of an instrument—a manometer—he developed that could measure atmospheric CO2 with unprecedented accuracy. He was diligent and meticulous in his work, almost psychotic in his pursuit of new measurements in new places. In a 1989 interview with Earl Droessler,

Revelle commented that “Keeling’s a peculiar guy…He wants to measure CO2 in his belly…” If his other measurements were any indication, in his belly the air would contain 315 parts per million of CO2. Gale Christianson provides perhaps the best, or at least the most entertaining, account of Keeling’s life and work

30 315 parts per million, but Keeling’s subsequent measurements over the course of the

1960s—and, in fact, the through the rest of the 20th century—would show that figure to be steadily rising. The “,” a saw-blade curve representing the gradual rise

of atmospheric CO2 at Mauna Loa since 1957, has since become an icon of global warming.41

Revelle’s reallocation of money from flexible grants made to Scripps by ONR and

AEC technically put Keeling’s atmospheric CO2 research on a government payroll, but it was his involvement in planning the International Geophysical Year that enabled Revelle

to put CO2 on the public research agenda. The IGY ran from July 1, 1957 to December

31, 1958, and was the largest cooperative international scientific research effort the world had ever seen. The idea for the IGY was born out of a gathering of physicists interested in the ionosphere at the house of James Van Allen in 1950; in less than a decade it blossomed into a project that involved more than 60,000 scientists and technicians from

66 nations.42 The initial group included Van Allen, a rocket scientist concerned primarily with cosmic rays; Lloyd Berkner, a physicists who would later become both the president of the ICSU and a member of Eisenhower’s Science Advisory Committee: Sidney

Champan, a man New York Times science writer Walter Sullivan dubbed the world’s

“greatest living geophysicist”; and three other physicists, J. Wallace Joyce, S. Fred

in a short section of Greenhouse, 151-157; see also Roger Revelle, interview by Earl Droessler, February 1989, American Institute of Physics, College Park, MD, as cited in Weart, Discovery, 36. 41 The “saw blade” character of the Keeling Curve represents the annual fluctuations in CO2 that result from the progression of the seasons in the Northern Hemisphere. With significantly more continental land mass than the Southern Hemisphere—and an accompanying deciduous biomass that sequesters carbon in the form of leaves and stems in the summer and then releases that carbon when those leaves die in the winter—the Northern Hemisphere dictates the annual biological “respiration” of global CO2. G.M. Woodwell, R.H. Whittaker, W.A. Reiners, G.E. Likens, C.C. Delwiche, and D.B. Botkin, “The Biota and the World Carbon Budget,” in Science, Vol. 199 (January 13, 1978), 141-146. 42 For the founding of IGY, see Water Sullivan, Assault on the Unknown: The International Geophysical Year (New York: McGraw Hill, 1961): 4-35; Fae L. Korsmo, “The Genesis of the International Geophysical Year,” Physics Today (July 2007): 38-43.

31 Singer, and Ernest H. Vestine.43 Envisioned as an updated version of the International

Polar Years of 1882-83 and 1932-33, the IGY had a major and important polar science component, but the project focused primarily on the physics of the Earth’s atmosphere.

In fact, with projects in meteorology, climatology, ionospheric physics, aurora and airglow, cosmic rays, and solar activity, the IGY agenda essentially defined what would soon become known more broadly as “atmospheric science.”44

Scientists designed the IGY to tackle a range of large-scale research interests, but from the beginning it was also a deeply political affair. Not surprisingly, in the international arena, the politics of the IGY revolved primarily around U.S.-Soviet relations (although the Antarctic components of the program also engendered tensions between other nations with political claims on that continent).45 Planning for the IGY was administered by the ICSU, with support from the United Nations Educational,

Scientific, and Cultural Organization (UNESCO) and the World Meteorological

Organization (WMO). Under Stalin, however, the USSR had decided not to adhere to the

ICSU (though the Soviets did belong to the International Astronomical Union and the

WMO), and the Soviet Union was left off of the original list of 26 invited nations. The

U.S. National Committee for the IGY had to persuade the ICSU to send the Soviets a special invitation to participate in the project.46 The Soviets took 18 months to accept; once they did, their cooperation was limited. Soviet officials resisted proposals to allow

43 Sullivan, Assault on the Unknown, 22; Korsmo, “The Genesis of the International Geophysical Year,” 40. 44 Thomas Malone recalls a fierce battle over the changing of the name of the Journal of Meteorology to the Journal of Atmospheric Science in 1960 during his term as president of the American meteorological society. The term, Malone remembers, came from AAAS president Paul Klopsteg. “It caused great pain to Werner Baum when we did that,” he told Early Droesser in 1989. Thomas F. Malone, interview by Earl Droessler, February 18, 1989, American Meteorological Society/University Corporation for Atmospheric Research Tape Recorded Interview Project. 45 Sullivan, Assault, 412, 415. 46 Korsmo, “Genesis,” 41.

32 scientifically-oriented over-flights by foreign nationals within Soviet borders, and Soviet scientists refused to discuss their nation’s rocketry and artificial programs. The rest of the Soviet IGY program was every bit as extensive as the United States’ proposed plan—perhaps more so—but cooperation between the two superpowers in sharing scientific resources and exchanging data took a back seat to what IGY boosters spun as healthy scientific competition between the communist East and the capitalist West.47 The recent detonation of hydrogen bombs by both sides—the Americans’ “Iron Mike” in

1952, the Soviets’ “Joe 4” in 1953—lent a certain urgency to this competition.48

Domestically, popular concerns about nuclear weapons testing and the perceived gap developing between the quality of American and Soviet science gave American scientific leaders like Revelle, Berkner, and Weather Bureau chief Harry Wexler the latitude to push for a comprehensive program of geophysical research for the

IGY—including research on CO2 and climate. Funding for the American portion of the

IGY came from Congress, primarily via the National Science Foundation. In May of

1957, members of the U.S. Committee on the IGY—Wexler, Berkner, and Revelle among them—went before the House Appropriations Committee to ask Congress for $39 million to run their 18 month program, including more than $2 million for oceanographic research and almost $3 million for research in meteorology.49 In wide-ranging testimony that covered everything from the aridity of the Martian atmosphere to the depth of the

47 The Soviets offered to provide nearly a third of the ships used for the IGY, and in 1955 they unveiled plans for three new permanent seismic stations in the Arctic, from where they could study all aspects of the crysophere. Korsmo, “Genesis,” 41. 48 There is a wealth of literature on the development of the atomic bomb, but for an introduction to the bomb and the scientists who made it, see Gregg Herken, The Brotherhood of the Bomb: The Tangled Lives and Loyalties of Robert Oppenheimer, Ernest Lawrence, and Edward Teller (New York: Henry Holt and Co., 2002); Kai Bird and Martin J. Sherwin, American Prometheus: The Triumph and Tragedy of J. Robert Oppenheimer (New York: Random House, 2005); Richard Rhodes, The Making of the Atomic Bomb (New York: Simon and Schuster, 1986). 49 U.S. Congress, Report on the International Geophysical Year, May 5, 1957, pg. 125.

33 Gulf Stream, Revelle introduced members of Congress to the possible relationship

between an increase in CO2 and an increase in global temperature, and noted the many uncertainties that a program in CO2 monitoring could hope to resolve. Congress approved the program, and ultimately IGY vessels and stations took CO2 measurements

50 at sixty locations around the world. For 18 months, CO2 stood on the front lines of Cold

War science.

Forecasting, GCMs, and Atmospheric Modeling

Revelle’s growing interest in CO2 ran parallel to another set of developments in climate-related atmospheric science also entangled in the web of Cold War research.

These were developments in atmospheric modeling, and in particular of General

Circulation Models, or GCMs. Early numerical models of the Earth’s atmosphere grew out of efforts to use computers to accurately forecast weather during and after the Second

World War, and in the ensuing decade scientists began to modify these models in order to attempt to predict changes in climate. Just as the nuclear weapons tests conducted between 1946 and 1953 allowed oceanographers to trace radioactive carbon through the oceans and atmosphere, so too did radioactive tracers provide modelers with important baseline data on how the gases of the atmosphere moved around the globe at various

altitudes and times of year. Even more so than CO2 monitoring efforts, models of the atmosphere had direct and obvious applicability to questions about nuclear fall-out and the potential distribution of radioactivity in the event of a nuclear exchange. In addition, some scientists and politicians drew a direct link between modeling the weather and

50 Sullivan, Assault, 240.

34 climate and controlling these geophysical forces, both for the domestic good via rainmaking, and as a powerful form of Cold War weapon. Ultimately, the combination

of the Scripps Institute’s CO2 research, the expansion of CO2 monitoring during the IGY, and this government–sponsored effort to create a realistic, predictive model of atmospheric circulation helped to pave the way for the institutionalization of atmospherically-oriented climate change research in America.

Denotatively, the relationship between weather and climate is relatively straightforward. “Weather,” the Oxford English Dictionary reveals, refers to “the condition of the atmosphere (at a given place and time) with respect to heat or cold, quantity of sunshine, presence or absence of rain, hail, snow, thunder, fog, etc., violence or gentleness of the winds.”51 The characteristic weather conditions of a country or region over time, in turn, constitute that region’s “climate.”52

The historical relationship between the disciplines of meteorology and climatology is more complicated. Despite the clear definitional relationships between the subject matter, until the 1950s, the prime objectives and key problems of climatology and meteorology were very different. The fundamental task of meteorology has almost always been to provide a more accurate and reliable prediction of the weather, and in the early 20th century, forecasting involved as much art as it did science. Weather

51 “Weather, n.1” The Oxford English Dictionary, 2nd edition, 1989, OED Online (Oxford University Press, 2000. http://dictionary.oed.com/cgi/entry/50282078. 52 The OED contains many definitions for climate; this is excerpted from 2.a. The first definition, which speaks to the etymology of the word, refers to a “band or belt of the earth’s surface stretching from west to east and associated with specific parallels of latitude.” These bands or belts, however, often also described different prevailing weather conditions; this first definition of “climate” thus offers some etymological insight into the second definition. In fact, the word “climate,” as I discuss in Chapter 3, comes from the Greek klima, or inclination, as in the inclination of the sun as seen from points on the Earth…which varies, not coincidentally, by latitude. Denis Hartman’s Global Physical Climatology (San Diego: Academic Press, 1994) defines climate more simply, and this definition or one like it is the definition commonly used by contemporary scientists. To Hartman, climate is “the synthesis of weather in a particular region.” 377. See “climate, n.1” OED Online, Draft Revision, July 2010 (Oxford University Press, 2010).

35 forecasters—mostly amateurs—typically relied on local knowledge and local records to help them predict relatively small, short-term changes in local conditions for practical purposes. Climatologists, on the other hand, focused predominantly on reconstructing past climates, and on using universal geophysical principles to try to explain long-term changes in those climates. They were, by and large, professional scientists working in university departments of geology or geography, and their work had few immediate practical applications.53

As a scientific discipline, meteorology was relatively late to professionalize, but when it finally began to do so in the 1930s, the field and its methodology changed rapidly. Impetus for expanding American meteorology came from many quarters, including the American Meteorological Society, the Weather Bureau, the Navy, and the nascent American airline industry. The science, however, came primarily from Sweden.

In the early 20th Century, a group of scientists studying under Vilhelm Bjerknes and his son Jacob in Bergen, Sweden, began to address weather phenomena in physical terms.

They focused not just on traditional “weather types” and conditions on the ground, but also on the movements of large masses of air and ocean that drove cyclones, storms, and, ultimately, just about every weather event under the sun. Their approach came to be known as “dynamic meteorology.”54 In the late 1920s and 1930s, many of these “Bergen

School” scientists began to appear as visiting scholars in the United States, and their work laid the foundations for the professionalization of American meteorology.

53 Between 1920 and 1929, U.S. universities awarded only six doctoral degrees in meteorology; of these six dissertations, three covered climate, and were written by geographers. Only one, a dissertation on free-air pressure maps written by Clarence LeRoy Meisinger, had a direct link to weather. Harper, 56. 54 See Harper. For parallels in oceanography, also at the Bergen school, see Helen Rozwadowski, The Sea Knows No Boundaries: A Century of Marine Science Under ICES (Seattle: University of Washington Press, 2002) and Fathoming the Ocean: The Discovery and Exploration of the Deep Sea (Cambridge: Harvard University Press, 2005).

36 The advent of dynamic meteorology in the United States coincided with a sudden and dramatic increase in demand for accurate meteorological information and from the American military and its allies during WWII. Led by the Swede

Carl Gustav Rossby, by 1943 the five universities represented in the War Department’s

University Meteorological Committee—MIT, NYU, Caltech, UCLA, and the University of Chicago—had trained approximately 8,000 meteorologists and 20,000 meteorological observers, technicians, and staff.55 Their training focused on practical weather forecasting for the Army, Navy, and Air Force, and at war’s end, many of these military meteorologists found work within a rapidly expanding commercial aviation industry.56

This twenty-fold increase in meteorological professionals in America provided a critical mass of scientific personnel, and Rossby and his colleagues in academia used the discipline’s growth as leverage to influence the government’s post-war science agenda for the benefit of more theoretical meteorological research.57

For academic meteorologists, the key to understanding the atmosphere lay in numerical weather prediction; the key to numerical weather prediction, in turn, lay in computers. In order to describe the dynamic processes of atmospheric motion, scientists had to solve long series of nonlinear equations, and they had to solve them quickly.

During the First World War, British Meteorologist Lewis Fry Richardson surmised that it would take about 64,000 human computers solving equations continuously in order to keep up with the Earth’s weather as it unfolded—and that’s to say nothing of predicting future weather.58 During the Second World War, however, two electrical engineers

55 Harper, Weather by the Numbers, 7, 76-89.. 56 Ibid., 84-85. 57 Ibid., 7, 84-85. 58 Ibid, 2.

37 working for Army Ordnance, Presper Eckert and John Mauchly, devised a machine that helped begin to solve this computing problem, the Electronic Numerical Integrator and

Computer, or ENIAC.59 Recognized as the first digital computers, the ENIAC and its successor, ENVAC (Electronic Discrete Variable Automatic Computer), were designed to compute ballistic weapons firing tables more quickly, but they could also solve the differential equations that described meteorological phenomena.

Following the suggestion of physicist Vladimir Zworykin (a pioneer in television tubes), meteorologists like U.S. Weather Bureau chief and his eventual successor Harry Wexler soon began to push for a computer-generated numerical weather forecast.60 The result, after more than five years of research and development, was the interdisciplinary Joint Numerical Weather Prediction Unit. The JNWPU—later reorganized as the Weather Bureau’s Geophysical Fluid Dynamics Laboratory (GFDL) under Joseph Smagorinksi at Princeton’s Institute for Advanced Study—used mathematician John von Neumann’s new and improved computing machine, along with the expertise of meteorologists, physicists, and mathematicians from all over Europe and the United States, to produce the world’s first 12 and 24 hour computer assisted numerical weather forecast in March of 1950.61

In the mid ‘50s, as the IGY approached, scientists and their sponsors at the

Weather Bureau and the NSF began to look at models as a way to go beyond forecasting to a deeper understanding of the geophysical processes of the atmosphere. With 12 and

59 For more on ENIAC—and the controversy over its development and the eventual title of “father” of the digital computer—see Scott McCartney, ENIAC: The Triumphs and Tragedies of the World’s First Computer (New York: Walker, 1999); Alice Rowe Burks, Who Invented the Computer? The Legal Battle That Changed Computing History (New York: Prometheus Books, 2003); Eric G. Swedin and David L. Ferro, Computers: The Life Story of a Technology (Baltimore: Johns Hopkins University Press, 2007). 60 Harper, Weather by the Numbers, 96-97. For more on Zworykin, see Albert Abramson, Zworykin: Pioneer of Television, (Champaign, IL: University of Illinois Press, 1995) 61 Harper, Weather by the Numbers, 141.

38 24 hour forecasts continuing to improve, meteorologists like Jule Charney and Joseph

Smagorinski of the JNWPU, as well as Victor Starr of MIT, turned their attention to

“General Circulation Models,” mathematical representations not just of the weather, but of the movement of the atmosphere as a whole.62 In 1955, Smagorinski and Wexler launched a spin-off of the JNWPU focused specifically on GCMs. They called the unit the General Circulation Research Section (soon renamed the General Circulation

Research Laboratory and eventually folded into the GFDL in 1962).63 Based on the physical properties of the atmosphere as expressed through a set of equations for fluid dynamic motion called the Navier-Stokes equations, GCMs described the typical movements of the air masses and atmospheric gases that affected the weather throughout the globe.64

Smagorinski and his colleagues began to build GCMs primarily as a of their weather-related work, but from the beginning, their models had clear Cold War

applications. As with studies of CO2, GCMs benefited greatly from the massive quantities of radioisotopes dumped into the atmosphere by nuclear weapons tests at

known times and places. General circulation modelers traced radioactive CO2 and other irradiated material through the atmosphere to provide observational data for their

62 Edward Lorenz, “Reflections on the Conception, Birth, and Childhood of Numerical Weather Prediction,” Annual Review of Earth and Planetary Sciences, Vol. 34 (May 2006): 37-45. Here, Lorenz mostly writes about the roles of Norman Phillips and Joseph Smagorinski, but his story includes most of the major players at the JNWPU, including Starr, who worked on the project while based at MIT. In a dialogue recorded in 1986, Phil Thompson and Ed Lorenz both briefly recall working with Starr on his GCM project at MIT, though they are unfortunately short on details about their mentor. “Dialogue between Phil Thompson and Ed Lorenz,” July 31, 1986, AMS/UCAR Tape Recorded Interview Project. 63 Interview with Joseph Smagorinski by John Young, May 16, 1986, AMS/UCAR Tape Recorded Interview Project, (pg. 20). 64 G.K. Batchelor, An Introduction to Fluid Dynamics (Cambridge: Cambridge University Press, 1967), 147.

39 models.65 Their models, in turn, provided a way to predict the distribution of radiation and fallout, as well as a method for detecting and even locating foreign nuclear testing activity as the resulting fallout scattered to the now-predictable four winds.66

Many meteorologists, alongside their colleagues in other geophysical sciences, sought an even more complete picture of the engines of geophysical motion that began with a model of the Earth’s radiation budget. For forecasting weather, computer modelers used what were essentially static, best-guess estimations of long-term geophysical processes to isolate the physical properties of land, water, and air that had an immediate impact on temperature, humidity, wind speed, and precipitation. But as meteorologists knew, some of the very factors that numerical weather models filtered out in their initial conditions—heat and wind driven ocean circulation, for example, or the chemical composition of the atmosphere—had important impacts on weather patterns and overall atmospheric circulation.67 Ultimately the distribution of energy—the Earth’s radiation or heat budget—drove the whole show. In the 1950s, scientists began to draw on theoretical, mathematic, and even tangible physical models of the Earth’s radiation budget to understand the weather as it changed over long periods of time: that is, to understand the climate.68

65 Hart and Victor, “Scientific Elites,” 648. 66 As recalled in an interview in 1993, the Weather Bureau had been working with the Air Force since the late 1940s to help detect Soviet nuclear bomb tests based on radioactive debris moving through the atmosphere. Essentially, as Machta described, the Weather Bureau used intercepted Russian meteorological data, combined with their own knowledge, to determine the most likely places to pick up radioactivity in the event of a detonation based on prevailing weather patterns. The Air Force then used Weather Bureau recommendations to decide where to send reconnaissance flights equipped with filters to collect radioactive particulate matter. Interview with Lester Machta by Julius London, October 31, 1993, AMS/UCAR Tape Recorded Interview Project (pg. 5). 67 For JNWPU’s (and later GFDL’s) attempts to integrate these problems in the 1960s, see Harper, Weather by the Numbers, 233. 68 Dave Fultz’s “dishpan” model of the atmosphere represented the far opposite end of the modeling spectrum from GCMs. Remembering that the atmosphere was in fact a fluid spinning around a point (a line, really, of diameter), Fultz devised simulations of atmospheric motion in spinning dishpans.

40 Masters of Infinity: Weather, Climate, and the Ethos of Control

Embedded in the various government-sponsored efforts to model the Earth’s atmosphere and forecast its weather was an optimistic idea that if humans could better understand the vagaries of atmospheric processes, they could learn to control them.

Pioneered in 1946 by General Electric’s Nobel Prize winning chemist Irving Langmiur and his colleagues Bernard Vonnegut and Vincent Schaefer, small and medium scale weather control projects appealed to both military and civilian agencies in the early Cold

War. The discourse on weather and climate control quickly grew to involve both domestic interests and foreign policy concerns.

Domestically, weather control projects promised to tackle real problems and, like most government projects, to bring money to the states they were conducted in. Boosters claimed that Langmuir’s method of cloud seeding might disperse fog at airports, help to alleviate droughts, or redirect dangerous storms—all of which appealed to Congressmen, especially those from the arid American West and the hurricane-prone Southeast. In the

1940s and 1950s, most of the applied weather modification research in the U.S. was conducted with small scale, domestic projects in mind.69 In 1953, Congress created an

Advisory Committee on Weather Control to oversee these projects.70

Meteorologists looked on these with some interest; in November of 1956, after a dinner at Chicago’s Steak and Saddle Room and a subsequent day full of meetings, the NAS Committee on Meteorology—which included Wexler, Revelle, and Berkner, among others—broke up to take a tour of Fultz’s lab. They were duly impressed. Minutes, National Academy of Sciences Committee on Meteorology, , November 27-28, 1956, Roger Revelle Papers (MC6), Box 28, Folder 7, Scripps Institute of Oceanography Archives, La Jolla, CA. 69 The most famous—or perhaps most notorious—of the domestic weather modification projects in the U.S. took place in South Dakota in 1972, when a flash flood in Rapid City followed shortly after a government sponsored rainmaking experiment. Nearly 200 people died in the flood, and the government came under fire from a class action lawsuit. Spencer Weart, “Climate Modification Schemes,” in The Discovery of

41 The military—and especially the air force—was also interested in local and regional weather modification. The same techniques that might benefit civilians at home might be put to effective military uses. Fog dispersal could help create an all weather air force, for example. So too could rainmaking slow an enemy’s communications and supply routes.71 More insidiously, the same techniques used to avoid drought at home could be used to create it abroad, wreaking havoc on an enemy’s food supply—a great asset in a non-shooting war.72 The possibilities seemed endless, for as Langmuir took pains to point out, the energy available in cumulous clouds exceeded that of even an atomic bomb.73

In addition to these local and regional applications of weather modification techniques, some scientists and politicians saw important military and foreign policy implications in potential weather modification projects conducted at a larger scale. They framed their concerns in terms of geophysical warfare more generally. In a paper presented to Congress in 1958, retired Navy Captain Howard T. Orville enumerated the ways in which humans might intentionally or unintentionally alter the weather or climate.74 Coincidentally, one of his first points involved the unintentional warming of

the Earth through CO2. CO2 was relatively benign compared to the possibilities

Global Warming: A Hypertext History of how Scientists Came to (Partly)Understand What People are Doing to Cause Climate Change, http://www.aip.org/history/climate/RainMake.htm. 70 U.S. Congress, “Weather Modification Research,” Hearing, House Committee on Interstate and Foreign Commerce, 85th Congress, 2nd Session, March 18-19, 1958: 9. 71 For more on the military applications of weather modification, see chapter 3. 72 James R. Fleming, “The Climate Engineers: Playing God to Save the Planet,” The Wilson Quarterly, (Spring 2007): 55. 73 Fleming, “Climate Engineers,” 53. 74 Howard Orville was Eisenhower’s closest advisor on weather matters, and the paper was hardly his first on weather modification. The idea had been on military minds since before Langmuire and his employees at General Electric began their experiments in 1946, and Orville published a piece in Collier’s in 1954 that contained possibilities for weather modification with explicit military applications. Fleming, “Climate Engineers,” 55.

42 introduced by intentional weather modification, however. The Russians might spread colored pigments over the poles in order to absorb solar energy and ultimately melt the ice caps, causing a drastic change in the local and even global climate. They might dam the Bering Strait and use plants to pump warm Pacific Ocean water into the Arctic Ocean, again, melting the ice caps to provide Russia with warm water ports at the expense of American coastal cities, which would be inundated. Finally—and perhaps worst of all—Orville cited a German Dr. Hermann Oberth, who foresaw “a gigantic mirror ‘hung’ in space” that would “focus the sun’s rays as a giant magnifying glass at any desired intensity and beam” in order to protect orchards from frost, melt ice in ports, or serve any number of less innocent climatic purposes.75

Orville’s paper—and the broader concern over weather and climate modification—reflected a great anxiety over Russian scientific superiority in the wake of the 1957 Sputnik launch, and government science boosters took advantage of this anxiety. As the always-hawkish nuclear physicist Edward Teller warned the Senate

Subcommittee on Military Preparedness in 1957,

“the Russians can conquer us without fighting through a growing scientific and

technological preponderance. Imagine, for instance, a world in which the

Russians can control weather in a big scale where they can change the rainfall

over Russia, and that—and here I am talking about a very definite situation—that

might very well influence the rainfall in our country in an adverse manner…What

75 Howard T. Orville, “The Impact of Weather Control on the Cold War,” in U.S. Congress, “Weather Modification Research,” 51-53.

43 kind of a world will it be where they have this new kind of control, and we do

not?”76

With Sputnik still fresh in Americans’ minds, Democrats like Lyndon Johnson began to attack what they saw as Eisenhower’s insufficient support for American science.

Research into weather and climate control provided a prime example of an area where the

U.S. must not get caught sleeping. “From space,” declared Johnson at the Democratic

Caucus in January of 1958, warning of the technologically and scientifically superior

Russians,

“the masters of infinity could have the power to control the earth’s weather, to

cause drought and flood, to change the tides and raise the level of the sea, to

divert the Gulf Stream, and change temperate climates to frigid.”77

He advocated more Congressional funding, both for weather modification research specifically and for science in general.

The move both bolstered Johnson’s image as a Cold Warrior and champion of

American science and pleased his colleagues from western states, including Republicans like Utah’s Aldous Dixon. Dixon, from one of the nation’s most arid states, testified in favor of weather control mostly on account of the promise of relieving droughts, but he concluded his testimony with a more patriotic appeal, referring explicitly to the Russian

76 Teller’s testimony was quoted more than once in the 1958 hearing, but interestingly, it never appeared the same way twice. U.S. Congress, “Weather Modification Research,” 1958: 53. Teller’s commitment to using nuclear weapons for both peaceful and warmaking purposes was no secret. A popular joke among physicists at the time ran, “You got a problem? Eddie’s got a bomb.” Zuoyue Wang, In Sputnik’s Shadow: The President’s Science Advisory Committee and Cold War America (New Brunswick, NJ: Rutgers University Press, 2008), 4. 77 Cited in U.S. Congress, “Weather Modification Research,” 1958: 52.

44 satellite. “There are many calls upon Congress for projects to meet the needs of the post- sputnik world,” Dixon noted.78 Weather control must be one of Congress’s priorities.

“America could become as subject to Russia’s whims as a rat in a laboratory to an

experimenter. If Russia beats us to the punch on learning how to control the

natural laws governing weather changes, she could conceivably produce a drought

over our whole continent or a disastrous flood. We know that the Russians are

devoting great energy and scientific talent to learning how to control the weather.

It is urgent that the United States not fall behind in this race.”79

Thus did weather and climate modification research, alongside forecasting, general

circulation modeling, atmospheric monitoring, and CO2, serve as a key points on the developing agenda of Cold War atmospheric science in America.

Big Science, Broad Science: The National Center for Atmospheric Research

In the 1950s, entrepreneurial scientists like Roger Revelle, Lloyd Berkner, and

Harry Wexler capitalized on a spike in government interest in geophysical science in order to garner financial support for a wide range of individual projects in atmospheric science. At the same time, atmospheric scientists and science administrators also set about creating permanent institutions to carry out the broad agenda of atmospheric research articulated during the planning of the IGY. Like scientists’ individual projects and like the IGY itself, these institutions were children of the Cold War. Domestically,

78 U.S. Congress, “Weather Modification Research,” 1958: 21. 79 Orville, Johnson, Teller, and Dixon appear to have been speaking hyperbolically of the Soviets’ plans to master the weather and climate, but their claim that the Russians were working on these things was actually not far off the mark. A rare Soviet book, Man vs. Climate, appeared in English sometime around 1961. In it, the authors detailed Russian scientists’ work on weather and climate modification, which was, as these boosters of weather modification research at home claimed, quite extensive. Rusin, N. and L. Flit. Man Versus Climate. Translated by Dorian Rottenberg. Moscow: Peace Publishers, 1960. U.S. Congress, “Weather Modification Research,” 1958: 21.

45 the most important of these institutions was the National Center for Atmospheric

Research (NCAR), located in Boulder, Colorado.

In 1956, National Academy of Sciences President Detlev Bronk appointed a

National Research Council Committee on Meteorology to “consider and recommend means by which to increase our understanding and control of the atmosphere.”80 The committee included Berkner, Rossby, Charney, von Neumann, Teller, Revelle (in an informal advisory role), and a number of other luminaries of meteorology and Cold War science. The committee noted a shortage of training, personnel, and dedicated resources in the study of the atmosphere.81 Meteorological research was fragmented and the discipline was underprofessionalized: 90% of American meteorologists were government employees and the vast majority of these military men had no Ph.D.s.82 The nature of the subject matter, meanwhile, required a sophisticated interdisciplinary approach, often using expensive, large-scale equipment that was at the time unavailable to most serious scholars.83 In early 1958, with the IGY underway, the committee recommended that the government increase its overall financial support for basic atmospheric research, and that the NSF underwrite the establishment of a National Institute of Atmospheric Research

(NIAR, later changed to NCAR) to coordinate and conduct this research.84

80 The official institutional history of NCAR leaves Revelle’s name out, and it is possible that Revelle only became a member of the Committee on Oceanography later that same year. He was involved in subsequent organizational meetings, however, as both his papers and the “green book” show. From “Statement by D. W. Bronk,” 5 February 1958, in Preliminary Plans for a National Center for Atmospheric Research: Second Progress Report of the University Committee on Atmospheric Research, February, 1959, Appendix C, C-1. Available at http://www.ncar.ucar.edu/publications/pubsabout.php and hereafter known as the “Blue Book.” See also Elizabeth Lynn Hallgren, The University Corporation for Atmospheric Research and the National Center for Atmospheric Research, 1960-1970: An Institutional History [National Center for Atmospheric Research, 1974): 3 (hereafter known as the “Green Book”). 81 Green Book, 3. 82 Ibid, 4. 83 Ibid. 84 Ibid, 3.

46 The Committee on Meteorology recommended that the Federal government fund

NIAR, but they strongly suggested that the new institution be run by representatives from meteorology departments at American universities rather than by government bureaucrats. Their vision for a university-based research program reflected an ambivalence about the Cold War research system. On one hand, the Committee saw

NIAR as a way to take advantage of government money in order to address a constellation of scientific, civilian, and military concerns, all at least ostensibly in the service of the state. Members of the Committee on Meteorology recognized that only the federal government could provide the resources and facilities scientists needed to approach the “fundamental problems of the atmosphere on a scale commensurate with their global nature and importance.”85 Airplanes loaded with atmospheric monitoring devices, expensive ground-based instruments including powerful telescopes and coronographs, and access to new technologies like computers and outstripped the financial capabilities of even the most well-endowed individual universities. Funding for these types of facilities would have to come from the government research establishment—the Department of Defense, ONR, and the NSF—and NIAR would in that sense itself be a government institution.

On the other hand, atmospheric scientists feared the hierarchy, secrecy, disciplinary compartmentalization, and strict oversight that might accompany government support. In particular, they worried that that the military would come to dominate atmospheric research.86 By channeling government resources into an institution expressly created to serve the interests of the universities, atmospheric scientists could insulate themselves

85 Blue Book, 6. 86 Green Book, 5.

47 from the direct oversight of their government sponsors. In February of 1958, the

Committee on Meteorology began working closely with the University Corporation for

Atmospheric Research (UCAR), a group of leaders from the nation’s university meteorology departments, in order to establish the structure and goals of NIAR.87 In early 1959, UCAR officially became the parent organization of the new institution

(which had by then been renamed NCAR). The UCAR board of trustees submitted a detailed proposal to the NSF that outlined NCAR’s purpose, institutional structure, and initial research plans—a proposal published as NCAR’s founding document, the “Blue

Book.”88

The Blue Book described a unique variation on the “big science” projects that characterized government-sponsored research in other fields—especially high energy and particle physics—during the Cold War.89 UCAR certainly wanted NCAR to operate at a large scale; in fact, the global nature of the subject matter all but demanded the institution think big, and the Blue Book explicitly stated that NCAR should provide “facilities and

87 UCAR’s members formally announced their intention to incorporate in October of 1958, and they did so officially in the state of Delaware in March of 1959, though the organization had been operational for quite some time. Tom Malone, Secretary of the American Meteorological Society and Director of Research for the Traveler’s Insurance Company, served as executive director pro tem. Blue Book, 4; Green Book, 17- 22. 88 The Blue Book was originally the “Second Progress Report of the University Corporation for Atmospheric Research,” February, 1959, Archives of the National Center for Atmospheric Research. The NCAR Blue Book was not the only important research proposal for atmospheric science with that nickname; as Erik Conway points out in Atmospheric Science at NASA: A History (Baltimore, MD: Johns Hopkins University Press, 2008), a later proposal for the Global Atmospheric Research Program (GARP), penned mostly by Jule Charney in 1966, was also referred to as the “Blue Book” because of the color of its cover. National Academy of Sciences—National Research Council, Feasibility of a Global Observation and Analysis Experiment, Publication number 1290, 1966, cited in Conway, 67. 89 Coined in 1961 by Alvin Weinberg of the Oak Ridge National Laboratory, “big science” is a general term that typically describes expensive, large-scale, government-sponsored research projects run either by private corporations or large scientific institutions. Weinberg coined the term in a cautionary article that was itself a response to President Eisenhower’s farewell address, in which he warned of the increasing power of the “military-industrial complex.” See Alvin Weinberg, “Impact of Large-Scale Science on the United States,” Science, 134 (1961),: 161-64; Alvin Weinberg, Reflections on Big Science (Cambridge: MIT Press, 1967). See also Peter Galison and Bruce Hevly (eds), Big Science: The Growth of Large-Scale Research (Stanford, CA: Stanford University Press, 1992).

48 technological assistance beyond those that can properly be made available at individual universities.”90 As Bruce Hevly notes in “Reflections on Big Science and Big History,” however, “big science” in the 20th Century was typically organized hierarchically, with vertically integrated teams striving for institutionally-directed scientific objectives.91

This was not the case at NCAR. If a typical big science project—the Berkeley cyclotron, for example—ran something like a business, where a CEO or board of directors dictated a set of research priorities with specific goals in mind and then hired scientists to do the job, NCAR operated more like a cooperative, where visiting scientists from member universities could have access to and control over facilities otherwise unavailable to them in a typical university setting. As an administrative necessity, the Blue Book divided the institution into three main sections along disciplinary lines—physical, mathematical, and chemical laboratories—but both these divisions and the hierarchies within them were intended to be flexible. The semi-permanent nature of NCAR’s scientific staff, composed largely of visiting scientists and postdoctoral fellows, along with a genial, well-liked, hands-off director in Walter Orr Roberts, further subverted the institution’s

90 Blue Book, vii. 91 Bruce Hevly, “Reflections on Big Science and Big History,” in Galison and Hevly, Big Science, 357. For more on “vertical integration,” see Robert Kargon, Stuart W. Leslie, and Erica Schoenberger, “Far Beyond Big Science: Science Regions and the Organization of Research and Development,” in Galison and Hevly, Big Science, 335. Kargon et all also talk about “horizontal integration,” a term they use to describe the relationships between centers of research and the businesses, universities, and other social elements of communities in which they are built. NCAR is an interesting case here, not least because of the evolution of the physical geography of the institution in Boulder. The famous main NCAR building, the Mesa Lab designed by I.M. Pei, is an iconic structure physically removed from—in fact put on high overlooking—the city of Boulder and the University of Colorado that sits at the city’s center. A ride on the NCAR shuttle, however, reveals that NCAR maintains research facilities throughout Boulder, and as it has grown, the institution has become more and more deeply interwoven into the town. It has also become part of a research community that includes the National Oceanographic and Atmospheric Administration, the University of Colorado, and a number of private and non-profit organizations with varying degrees of university affiliation like the Center for Severe Weather Research, the Natural Hazards Center, and the National Snow and Ice Data Center.

49 administrative and scientific hierarchies.92 “The goal,” UCAR’s trustees argued, “is to achieve the fundamental unity of the atmospheric sciences while respecting the identity of the basic disciplines which have a part to play in atmospheric research.”93

The scale and complexity of the atmosphere itself added to the singularity of the

“big” atmospheric science envisioned for NCAR. Most big science projects in America revolved around an object or single, tangible objective—the cyclotron, the atomic bomb, the gravity probe, or the space telescope. These projects’ managers typically approached complex systems with relatively single-minded goals—to collide particles, to collect astronomical data—and they tended to treat their projects and experiments in isolation from the forces of the outside world. By contrast, atmospheric scientists launched NCAR as a way to study the composition, motion, and behavior of the various elements of a global space that in fact was the outside world. UCAR’s board of trustees presented

NCAR as a sort of base station in a much larger effort to study the atmosphere at its full, global scale.94 Simply describing the atmosphere—let alone predicting or controlling it, as atmospheric scientists hoped to be able to do—required not only the flight facilities and measurement and observation instruments available to NCAR locally, but also access to a vast array of external data from monitoring stations, and later satellites, around the globe.95 Atmospheric scientists used NCAR’s computers to generate radiation budget

92 See Walter Orr Roberts, “Early History of the High Altitude Observatory,” an oral history conducted by the UCAR oral History Project, July 5, 1966, UCAR/NCAR Archives; David H. DeVorkin interview of Walter Orr Roberts, July 26-28, 1983, Oral History Program, National Air and Space Museum of the Smithsonian Institution. 93 Blue Book, 49-50. 94 Green Book, 46, citing the NCAR 10-year plan. 95 Atmospheric scientists were very clear about the relationship between description, understanding, prediction, and control in the Blue Book. “For the atmosphere,” they wrote, “as for any physical system, the capacity to understand is predicated on the capability to describe; the ability to predict in a really satisfactory manner is dependent on the capacity to understand and in a scientific sense is a measure of that understanding; intelligent control must rest firmly on an ability to predict—else the efficacy of measures by

50 models and GCMs, for example, but they could only do so using data generated elsewhere (mostly during the IGY). NCAR was intended as “an international intellectual center with a marked interdisciplinary flavor” capable of serving as a focal point in a developing field of science that was not only “big” in personnel and resources, but also geographically and disciplinarily broad.

NCAR began operations in Boulder in 1960, and when it did it marked an institutional confluence of four important and overlapping strands of atmospherically- oriented climate research: radiation budget modeling, general circulation modeling,

research on weather and climate control, and the CO2 question. For the UCAR board, an integrated model of the energy processes of the atmosphere was ripe for the picking. “In the first place,” the board argued in the Blue Book, “the development of mathematical models of atmospheric motion has now reached the stage at which it is feasible to introduce the energy equation in explicit form.”96 Introducing elements from radiation models—mathematical descriptions of the flow of energy from the sun, including the distribution and flux of particle energy, short and long wave radiation, and heat exchanges between the atmosphere and the Earth’s surface, through an otherwise static atmosphere—severely complicated models of circulation.97 In order to incorporate radiation equations into GCMs effectively, atmospheric scientists needed more information on world-wide sources and sinks of solar energy. Fortunately, as the Blue

Book pointed out, “the imminent advent of meteorological satellites will place in the hands of the atmospheric scientist a new and remarkably powerful tool for viewing the

which control is attempted will remain in doubt because of uncertainty as to the course the atmosphere could have taken if left to its own devices.” Blue Book, 32. 96 Blue Book, 38. 97 Ibid., 37.

51 heat balance problem in its global entirety.”98 Another new Cold War institution, the

National Aeronautics and Space Administration (NASA, created in 1958), was already hard at work on these satellites.99 To atmospheric scientists, this kind of fundamental knowledge of the Earth’s energy budget represented the keystone of any effective effort to control the processes of the atmosphere. The Blue Book stated the case plainly. “The physical linkage between the heat budget and the general circulation of the atmosphere is such a close one that any hope of effective climate control is likely to lie in alteration of some aspect of the heat budget.”100

The Blue Book expressed atmospheric scientists’ optimism about the possibilities of eventually controlling the climate, but it also revealed the seeds of concern about

unintentional climate modification, particularly through the accumulation of CO2.

Atmospheric scientists identified atmospheric CO2 as one of a number of poorly understood physical and chemical phenomena that could complicate their models and understanding of the global atmosphere. As they did elsewhere, in the Blue Book these

scientific leaders associated questions about CO2 with questions about nuclear testing.

“Man's activities in consuming fossil fuels during the past hundred years, and in

detonating nuclear weapons during the past decade have been on a scale sufficient

to make it worthwhile to examine the effects these activities have had upon the

atmosphere. Reference is made here to the still unsolved question of whether the

carbon dioxide content of the atmosphere is increasing as a result of combustion

98 Ibid., 39. 99 NASA, too, became interested in atmospheric modeling in the 1960s, and by the end of the decade, NASA’s Goddard Institute of Space Studies had become one of the international centers of both global circulation modeling and climate modeling. For more on atmospheric science at NASA, see Conway, Atmospheric Science at NASA. Thanks to Erik for providing me with a copy of his unpublished manuscript for use on the early stages of this dissertation. 100 Blue Book, 39.

52 processes and the even more elusive question as to possible changes in the earth's

electrical field as a result of nuclear explosions.”101

The plan to study CO2 and climate at NCAR hardly represented a clarion call on global warming; it was, like Revelle’s famous “geophysical experiment” comment, primarily a statement of scientific uncertainty. Nevertheless, atmospheric scientists, attuned to the

dangers of other forms of geophysical modification, sowed their doubts about CO2—and about Cold War science more generally—into the institutional soil of atmospheric science.

CO2 as Pollution

In the 1960s, these seeds of doubt began to germinate. Since as early as 1957,

when Revelle testified before Congress on CO2 monitoring in the IGY program, the

handful of atmospheric scientists interested in CO2 monitoring had used the potentially disruptive impacts of CO2 on climate as a way to lobby for financial support. A rise in

global CO2 could raise temperatures, they argued, which in turn could melt glaciers, leading to a rise in the sea level. Changes in temperature might also lead to changes in weather that would affect agricultural output in certain regions—for the better in some places, for the worse in others.102 Making his case before Western Congressmen in 1957, for example, Revelle suggested that minor changes in the Earth’s climate could threaten the water supply of the American Southwest.103 As Revelle and other scientists made

101 Blue Book, 45. 102 Sea-level rise had been associated with warming at least as early as 1953, when , citing Gilbert Plass and G.S. Callendar, among others, speculated that continued warming might lead to the inundation of American cities. Leonard Engel, “The Weather is Really Changing,” New York Times Magazine, July 12, 1953. 103 U.S. Congress, Report on the International Geophysical Year, May 5, 1957: 106.

53 sure to point out, though, these were mere conjectures—vague possibilities whose likelihood remained unknown but whose potential consequences made the curious

question of CO2 accumulation a question worth studying. “Only God knows whether what I am saying is true,” Revelle admitted to his Congressional audience. “What I am driving at is that this business of the carbon dioxide production is in fact a way of studying climatic changes.”104

In the late 1950s and early 1960s, scientists began to chip away at this uncertainty,

especially as it applied to the basic accumulation of CO2. In 1958, Swedish meteorologists Bert Bolin and Erik Eriksson helped to clarify the nature of the exchange of carbon dioxide between the oceans and the atmosphere that Revelle and Suess had discussed briefly in their now famous Tellus article of the previous year. Based on a short-term model of atmosphere-ocean interactions that included a more sophisticated account of the buffering mechanism Revelle had referred to, Bolin and Eriksson

predicted that atmospheric CO2 would most likely increase by 25% by the end of the century.105 A re-calibrated measurement of the “Suess Effect” (the name given to the

dilution of C14 rich atmospheric CO2 by C14 depleted CO2 from the burning of fossil

fuels since 1850), alongside Keeling’s steadily-rising CO2 curve from his measurements at the Mauna Loa Observatory, seemed to confirm that atmospheric CO2, as Callendar had suggested years before, was, in fact, rising. 106 It also appeared that humans were the

104 Ibid., 108. 105 Bert Bolin and Erik Eriksson (1958), “Changes in the Carbon Dioxide Content of the Atmosphere and Sea due to Fossil Fuel Combustion,” in Bert Bolin (ed.), The Atmosphere and the Sea in Motion: Scientific Contributions to the Rossby Memorial Volume (New York: Rockefeller Institute Press, 1959): 130–142. Available alongside a concise interpretive essay and discussion questions through the National Digital Science Library at http://wiki.nsdl.org/index.php/PALE:ClassicArticles/GlobalWarming/Article8. 106 As the NDSL essay briefly suggests, Bolin and Eriksson found inaccuracies in the measured decrease in C14/C12 ratio identified by Seuss. Again, the reason for the inaccuracies involved the buffering mechanism. Because the oceans take up atmospheric CO2 slowly and on a limited basis, an increase in

54 primary cause.

Developments in CO2 research dovetailed with the broader scientific community’s increased awareness of and concern about related environmental issues. In the late 1950s and early 1960s, scientists from a variety of fields began to use their expertise to study the human and environmental impacts of the great scientific and technological achievements of the 20th Century. Marine biologist Rachel Carson’s influential 1962 exposé of the environmental and public health impacts of chemical pesticides, Silent

Spring, stands out as the most famous example of scientists’ renewed role in addressing environmental change in the 1960s.107 Revelle’s career path was perhaps more characteristic of most geophysical scientists’ involvement in environmental issues.108 His chief environmental concerns involved nuclear radiation, population, and pollution.

Building on his earlier work with the Navy, in the mid 1950s Revelle chaired an NAS committee on the Biological Effects of Atomic Radiation, which released a report to the public in 1956 (the so-called BEAR report).109 In 1961, he left Scripps to work as science advisor for Secretary of the Interior Stewart Udall, perhaps the most vocal supporter of

atmospheric CO2 from fossil fuels with a low C14/C12 ratio will be accompanied by a much smaller percentage increase in oceanic CO2 of the same C14/C12 ratio. Because of the existing C14 rich CO2 in the ocean, ratio of C14/C12 in oceanic CO2 and atmospheric CO2 will be different; “the C14 between the two reservoirs would no longer be in equilibrium,” and C14 would transfer from ocean to atmosphere, masking the dilution of pre-existing, radio-carbon rich atmospheric CO2 by new, radio-carbon poor CO2 from fossil fuels. Bolin and Eriksson, 131. 107 Rachel Carson, Silent Spring (Boston: Houghton Mifflin, 1962). Carson herself followed in a long tradition of American scientists speaking out against environmental degradation, beginning as early as George Perkins Marsh’s 1865 Man and Nature [2003 ed], ed by David Lowenthal (Seattle: University of Washington Press, 2003). Among many others, see also William Leroy Thomas (ed.), Man’s Role in Changing the Face of the Earth (Chicago: University of Chicago Press, 1956); Paul Sears (1937) Deserts on the March [4th Edition] (Norman: University of Oklahoma Press, 1980). 108 It is likely that Revelle knew of Carson beforehand from her previous book, The Sea Around Us (New York: New American Library, 1954), which applied more directly to Revelle’s field. In any case, Revelle remembered discussing Silent Spring with Udall in an interview conducted by a television station in San Diego in 1992. Roger Revelle: Statesman of Science, a production of KPBS Television, San Diego, taped August 17, 1992, transcript, p. 19, SIO Archives, 92-39, UCSD. The interview is cited in Day, Revelle Biography (not paginated). 109 Day, Revelle Biography (not paginated).

55 conservation in the Kennedy and Johnson Administrations and the man largely responsible for putting the Great Swamp National Wildlife Refuge in New Jersey, the

Clean Air Act of 1963, the Wilderness Act of 1964, and the Endangered Species Act of

1966 on the Presidential agenda.110 In 1964, Revelle moved to Cambridge to head

Harvard’s new Center for Population Studies, where he focused on the “consequences of population change on human lives and societies, and of the biological, cultural, and economic forces that influence human fertility.”111 Revelle was not necessarily an environmentalist; rather, he shared with his colleagues a commitment to the sensible scientific management of the environment and natural resources for the benefit of humankind.112

Many atmospheric scientists echoed Revelle’s personal and professional interest in

environmental issues, and in the early 1960s they began to address CO2 in terms of a buzzword of the environmental movement: pollution. Initially, these scientists drew a

loose association between CO2 research and air pollution, much in the same way that they

had drawn an association between CO2 and radiation since the mid 1950s. For example, in a 1962 NAS-NRC report on The Atmospheric Sciences, 1961-1971, commissioned by

President Kennedy’s science advisor Jerome Wiesner, the Committee on Atmospheric

110 Udall passed away during the writing of this dissertation, and a flowering of brief memoirs and obituaries have poured forth in the media, but a critical biography of Udall has yet to be written. Boyd L. Finch’s Legacies of Camelot: Stewart and Lee Udall, American Culture, and the Arts (Norman: University of Oklahoma Press, 2008) addresses the brothers’ role in the Kennedy cabinet through the lens of culture, but Stewart Udall—and the Udall family—have played a major role in three generations of American politics, and their story awaits further historical exploration. 111 “Revelle, Roger,” in McGraw-Hill Modern Scientists and Engineers, (New York: McGraw Hill, 1980), as cited in Day, Roger Revelle (not paginated). 112 Reflecting on Revelle’s career, Walter Munk took pains to note that “Roger Revelle advocated the sensible use of the environment for the benefit of human beings, but did not ally himself with conservationist and environmental societies.” Munk’s insistance that Revelle kept a distance from environmentalists speaks to a persistent wariness on the part of scientists of political activism—and especially of environmental activism, as I will discuss in later chapters. Day, Roger Revelle (not paginated).

56 Sciences proposed to study water vapor and CO2 alongside various other trace gases—ozone, methane, and oxides of nitrogen and sulfur, and “several radioactive gases”—associated with industrial air pollution. “All these substances are important,” the Committee wrote.

“Some act as nuclei that set of condensation and precipitation; others influence the

energy transformations; some are radioactive; and many of them are of interest in

connection with air pollution.”113

The relationships between these gases and their potential impacts remained vague, but the

report implied that CO2 accumulation, inadvertent weather modification, atmospheric radiation, and air pollution were all of a piece.114

As a pollutant, however, CO2 was problematic. A report from a Conservation

Foundation conference headed by Charles David Keeling in 1963 articulated the problem.

CO2 provided a very good indicator for smog, haze, and other forms of pollution associated with fossil fuel consumption, but Keeling noted that “air pollution in the

115 ordinary sense does not include the CO2 rise in the atmosphere.” Whereas substances like DDT, Strontium 90, or sulfur dioxide directly threatened the health and wellbeing of

ecosystems and human populations, CO2 was a natural product of animal respiration and a gas essential to photosynthesis—and, by extension, essential to life on Earth. Both the

Air Pollution Control Act of 1955 and the Clean Air Act of 1963 dealt primarily with

113 The Atmospheric Sciences, 1961-1971: Volume I: Goals and Plans, Committee on Atmospheric Sciences, Publication 946 (Washington, D.C.: National Academy of Sciences-National Research Council, 1962), 33. 114 Ibid. 115 Implications of Rising Carbon Dioxide Content of the Atmosphere, (New York: The Conservation Foundation, 1963), 14.

57 toxic or noxious substances that had direct, typically local effects on human health.116

These substances typically emerged from discernable local point sources—cars or

factories—that government officials could regulate. CO2, by contrast, was harmless at the local level. Only through a complex series of geophysical processes resulting from a

global accumulation of the gas could CO2 impact human populations—hardly a phenomenon that state or federal governments were in a position to deal with.117

Nevertheless, Keeling saw important physical and philosophical connections

between the increasingly popular concern over pollution and the scientific interest in CO2

accumulation. Increased atmospheric CO2, the Conservation Foundation report argued, might ultimately act much in the same way as pollution, albeit through a more complex

chain of causation. CO2 could threaten human welfare through a rise in sea level, and might cause significant ecological and agricultural disruption through a warming of the air and ocean sufficient to change patterns of plant growth and species distribution.118

CO2 interacted with other fossil fuel pollutants—particularly sulfur dioxide—in unpredictable ways, further increasing the unpredictability of humans’ impact on the atmosphere.119

The Conservation Foundation’s concern over CO2 reflected the same precautionary ethos that spawned the Foundation’s concern about pollution. Like pollution, Keeling’s

report suggested, the accumulation of CO2 highlighted the dangers of unchecked

116 Air Pollution Control Act of 1955, Pub.L. 84-159, ch. 360, 69 Stat. 322; U.S. Public Health Service, “Clean Air Act, December 17, 1963, as amended October 20, 1965 and October 15, 1966,” (Washington, D.C.: U.S. Government Printing Office, 1967). 117 Congress was hesitant even to regulate traditional air pollution on a national level. The 1955 Air Pollution Control Act provided for federal research into air pollution, but effectively deferred to the states for actual pollution regulations. Only with the Clean Air Act of 1963—and its various updates over the ensuing decades—did the Federal government take up the problem of regulating harmful emissions. See J. Fromson, “A History of Federal Air Pollution Control,” in Ohio State Law Journal 30(1969): 517. 118 Implications of Rising Carbon Dioxide, 6-7. 119 Ibid., 4.

58 technological and economic “progress.” “The potentially dangerous increase of CO2, due to the burning of fossil fuels, is only one example of the failure to consider the consequences of industrialization and economic development,” the Conservation

Foundation wrote.120

“Man’s ability to change the environment has increased greatly over the last sixty

years and is likely to continue to increase for some time to come. Even now it is

almost impossible to predict all of the consequences of man’s activities. It is

possible, however, to predict that there will be problems…”121

In 1965, the President’s Science Advisory Committee attempted to catalogue these problems—and potential solutions to them—in a single, comprehensive executive report,

Restoring the Quality of Our Environment.122 Commissioned by Donald Hornig,

President Johnson’s Special Assistant for Science and Technology, the PSAC Panel on

Pollution provided a definition of pollution sufficiently broad to encompass a loose constellation of general and specific issues including the public health impacts of air pollution, water pollution, and pesticides; the impact of human actions on other living organisms; the “impairment of water and soil resources”; the “polluting effects of

123 household detergents”; urban decay; and the climatic effects of CO2. “Environmental pollution,” the report read,

“is the unfavorable alteration of our surroundings, wholly or largely as a by-product

of man’s actions, through direct or indirect effects of changes in energy patterns,

radiation levels, chemical and physical constitution and abundances of organisms.

120 Ibid., 15. 121 Ibid. 122 Restoring the Quality of Our Environment: Report of the Pollution Panel, President’s Science Advisory Committee, (Washington, D.C.: The White House, 1965). 123 PSAC, Restoring, 3-9.

59 These changes may affect man directly, or through his supplies of water and of

agricultural and other biological products, his physical objects or possessions, or his

opportunities for recreation and appreciation of nature.”124

According to the PSAC authors, though it only indirectly affected humans’ well-being

through the mechanisms of climatic change, CO2—the “invisible pollutant”—nevertheless constituted a potentially dangerous by-product of advanced industrial society that needed to be monitored and potentially counteracted in the future.125

Despite their growing concerns about the accumulation of CO2 from fossil fuels, however, in the 1960s geophysical scientists remained characteristically optimistic about their ability to counteract the gas’s climatic impacts through various forms of intentional geophysical modification. Revelle served as the chair of PSAC’s carbon dioxide group, and as usual he pointed out the pressing need for better understanding of atmospheric phenomena—understanding that could lead to control.126 The first step was to build a more comprehensive model of the atmosphere—a model that Revelle predicted the

Weather Bureau would have up and running in less than two years.127 The understanding gleaned from these models would ultimately allow humans to make conscious changes to other processes affecting climate.128 He was particularly sanguine about plans to change

124 Ibid., 1. 125 Ibid., 112. 126 The group also included Keeling, Smagorinski, oceanographer Wallace Broecker, and geochemist Harmon Craig. As Bolin and Eriksson had in 1958, the PSAC carbon dioxide group warned that by the year 2000, the burning of coal, oil, and natural gas could lead to a 25% increase in atmospheric CO2. “This will modify the heat balance of the atmosphere to such an extent that marked changes in climate, not controllable through local or even national efforts, could occur.” PSAC, Restoring, 112. 127 PSAC, Restoring, 121. 128 Revelle had painted a similar picture of the CO2 problem before President Johnson’s Domestic Council a year earlier, and in 1964 he called for similarly bold action. “With the advance of science and technology,” he wrote, “our power to change nature has grown enormously both for good and for ill. …by gaining

60 the Earth’s albedo—that is, the amount of the sun’s energy reflected back into space by the surface of the Earth—by scattering reflective particulates over swaths of the ocean.129

“The climatic changes that may be produced by the increased CO2 content could be deleterious from the point of view of human beings,” Revelle’s group noted. “The possibilities of deliberately bringing about countervailing climatic changes therefore need to be thoroughly explored.”130 This Cold War ethos of control would stand in uneasy balance with a growing commitment to the ethos of precaution for the next two decades.

Conclusion

It would be difficult to overstate the impact of the Cold War on the development of

atmospheric science and CO2 research in the United States. In 1938, when Guy Stewart

Callendar articulated his remarkably modern CO2 theory of climate, only a handful of meteorologists and climatologists took note, and among this small group few took the notion of anthropogenic global change seriously. By 1965, atmospheric scientists had

incorporated CO2 research into both individual projects and institutional directives at the national and international levels, and CO2 had appeared at least twice on the agenda of

131 the President of the United States. In part, the increased visibility of CO2 and climate arose out of advances in science itself. Papers on the physics and chemistry of CO2 by greater understanding, we will be able to make conscious changes—to bring more water to deserts, to bring cooler summers and warmer winters to the Middle West and the Northeast. In thinking about how we can make our country a better place in which to live by changing our environment, we must not be afraid of big things that can be done only on a national or international scale. We must be sure to make more than little plans.” Joseph Fisher, Paul Freund, Margaret Mead, and Roger Revelle, “Notes Prepared by Working Group Five, White House Group on Domestic Affairs,” April 4, 1964, President’s Committee [White House Group on Domestic Affairs], File 42, Box 20, Roger Revelle Collection MC 6, Scripps Institute of Oceanography Archives, La Jolla, California. 129 PSAC, Restoring, 127. 130 Ibid. 131 PSAC was the second Presidential task force to whom Revelle had introduced the issue of CO2. The first was a subgroup of Johnson’s Domestic Council, which released a report in 1964, as noted previously.

61 Callendar, Gilbert Plass, Revelle and Seuss, and Bolin and Eriksson, for example, alongside better models of the general circulation and the Earth’s radiation budget developed by Charney, Smagorinski, and others, began to paint a clearer picture of the

relationship between CO2 and climate in the late ‘50s and early ‘60s.

These scientific developments dovetailed with a set of philosophical, financial, and structural changes in atmospheric science that were ultimately more important in putting

CO2 on the popular and political map, however. These changes were by and large products of the Cold War. The advent of nuclear weapons established humans as agents of geophysical change, giving credence to the concept of anthropogenic climate change and spawning a great increase in government and military interest in geophysical processes like weather and climate. Leaders in atmospheric and other geophysical sciences like Revelle, Wexler, Berkner, and Charney soon began to capitalize on the newly available resources that accompanied this increased interest by framing research

on CO2 and climate in terms of its potential geopolitical significance. These scientists

promoted atmospheric science and CO2 research both through individual projects like

Keeling’s Mana Loa monitoring stations and the global monitoring effort of the IGY, and through the founding of institutions of atmospheric science, most notably NCAR.

Ultimately, the heightened administrative concern about the climatic impacts of CO2 that arose in the early 1960s reflected the political success of these scientists in incorporating

research on CO2 and climate, on atmospheric modeling, and on weather and climate modification into the larger agenda of government sponsored Cold War science in the late 1950s.

Many atmospheric scientists hoped to tackle CO2 and atmospheric circulation on a

62 world-wide basis, but interest in atmospheric science was by no means exclusively global. The short and mid-tem goals of efforts to predict and control weather and climate appealed primarily to politicians from the arid West and hurricane-ridden Southeast who recognized the tangible local benefits that might arise from a better understanding of meteorological events. Still, even domestic concerns about weather and climate entwined with the scientific objectives of the Cold War. Scientific leaders interested in the global phenomena of the atmosphere used the potential regional impacts of these phenomena to gain support from Western and Southeastern senators for their research. Senators from these regions in turn cited a patriotic concern over the national security implications of weather and climate control—alongside the benefits of local weather modification projects—in order to win appropriations for further research in this area. Boosters of atmospheric science made sure to emphasize both of these potential applications of their

basic research, especially when it came to the nebulous issue of CO2 accumulation. The combination of individual politicians’ interest in regionally-specific weather modification projects and the government’s broader interest in the potential geopolitical impacts of atmospheric change helped men like Revelle, Berkner, and Wexler establish atmospheric

science in general—and research of CO2 and climate in particular—as a Cold War research priority for the 1960s.

At the outset of the decade, atmospheric scientists’ vision for their field was justifiably optimistic.132 Both their scientific and their administrative success, made possible largely by newly available government money, fed this optimism. Relatively

132 In Chaos: Making a New Science (New York: Viking, 1987), James Gleick noted that the atmospheric science projects founded in the 1960s sprang up during “years of unreal optimism about weather forecasts.” This optimism, he suggests, relied on “an idea that human society would free itself from weather’s turmoil and become its master instead of its victim.”(pg. 18). Gleick is cited in Conway, Atmospheric Science at NASA, 65.

63 rapid and continuous improvements in weather forecasting and GCMs, along with better and more complete theoretical descriptions of atmospheric motion, convinced many atmospheric scientists—especially meteorologists and atmospheric modelers—that the primary obstacles to the ability to predict and eventually control the weather and climate were the availability of accurate global data and their computers’ capacity to handle that data. New Cold War technologies—especially the satellite and the increasingly powerful computers developed in the late ‘50s and early ‘60s—promised to remove both of these potential roadblocks. Beginning with NCAR, leaders in atmospheric science sought to create more permanent institutional arrangements for scientists to access these new tools.

In the early 1960s, led by Jule Charney, American atmospheric scientists also sought to expand the geographical reach of their research by institutionalizing the system of global atmospheric monitoring begun during the IGY.133 Working with the International

Council of Scientific Unions and the World Meteorological Organization (both connected to the United Nations, but in different capacities), Charney began work on the Global

Atmospheric Research Programme (GARP), a comprehensive system of monitoring and data sharing endorsed by President Kennedy that finally came online at the end of the decade.134 Scientists continued to emphasize the uncertainties and outstanding questions of their research, but they saw answers on the horizon.

Atmospheric scientists’ general optimism about the future of their profession

133 Conway, 65; see also Norman Phillips, “Jule Gregory Charney, January 1, 1917-July 16, 1981,” National Academy of Sciences Biographical Memoirs, http://www.nasonline.org. 134 Proponents of projects like GARP invariably referred to President Kennedy’s 1961 State of the Union Address, in which he made a relatively minor, passing comment about the need for international collaboration in measuring and monitoring the global atmosphere. Kennedy broached the subject again at the United Nations in August of the same year, but again, his proposal was relatively vague and tangential to his larger goals. The reference appears periodically in literature on international studies of meteorology and climate throughout the 1960s and ‘70s, however, usually phrased as if the President had sought to make atmospheric modeling a priority on par with sending a man to the . See especially Thomas F. Malone, “Reflections on the Human Prospect,” http://humanprospect-post2.blogspot.com.

64 masked an undercurrent of anxiety about the structure of the Cold War research system that made their success possible, however. Like many of their colleagues in other disciplines, they worried about the centralization and militarization of scientific research in the service of the state.135 The character and administration of their institutions—particularly NCAR, but also later GARP and, to an extent, the Geophysical

Fluid Dynamics Laboratory—reflected their efforts to insulate atmospheric scientists from some of the very government agencies that provided financial support for their research. As a result, atmospheric science became a particular form “big science,” unique not only in its broad, global reach, but also in its flattened management hierarchies and its user-driven research agenda.

Heightened concerns over the accumulation of atmospheric CO2 in the early 1960s also reflected atmospheric scientists’ growing ambivalence about the social and environmental impacts of science itself. Philosophically, scientists like Revelle and

Keeling shared a general interest in the unintended consequences of human actions with members of America’s bourgeoning environmental movement, and the two communities’ concerns had common Cold War roots. In the early 1960s, atmospheric scientists even

began to frame CO2 in the terms of the dominant paradigm in contemporary environmental politics: the paradigm of pollution. But atmospheric CO2 was an unconventional pollutant, and the scientists who studied it maintained a faith in their discipline’s ability to study, understand, and ultimately solve environmental problems

135 In 1967, the American Association for the Advancement of Science, an organization often on the progressive side of science activism, identified secrecy in science as one of the most important issues facing the American scientific community in the 1960s. The 1960s saw a flourishing of organizations and movements of concerned and dissenting scientists, as I discuss briefly in Chapter 2. Minutes, AAAS Committee on Science in the Promotion of Human Welfare, May 14, 1967, Papers of the Committee on Science in the Promotion of Public Welfare, AAAS Archives, Washington, D.C.

65 that many of their environmentalist counterparts had begun to doubt. Nevertheless, the

precautionary ethos that led them to continue to pursue CO2 as an environmental pollutant put atmospheric scientists in loose conversation with the growing American environmental movement in the 1960s. Atmospheric scientists were hardly environmental activists in the mid-1960s, but by the end of the decade, the scientific, political, and epistemological pieces were in place for that to begin to change.

66 Chapter 2 The Supersonic Transport: Scientists, Environmentalists, and the Atmospheric Environment

On June 5, 1963, the Kennedy Administration announced a large-scale cooperative program between industry and government to build a commercial passenger aircraft that would travel faster than the speed of sound. The American Supersonic

Transport, or SST, had little to recommend it economically, but as a bold statement of

America’s commitment to superiority in aerospace technology during the Cold War, the project initially found support in Congress and among the public. Over the course of the

1960s, however, Kennedy-era military-industrial aerospace projects fell out of favor with an increasingly skeptical public, and support for the SST waned.136 Critics questioned the wisdom of the expensive program in light of increased spending on the Vietnam War and a slowdown in the growth of federal science budgets. Environmentalists, growing in influence in American politics in the 1960s, argued that the marginal benefits of the plane could not justify its fuel consumption, its noise, and the disruptive sonic boom it left in its wake.

Environmentalists’ concerns revolved around the threats SSTs posed to

American’s everyday quality of life, but as the debate over the Supersonic progressed, scientists also began to raise concerns about the plane’s impact on the global environment. As scientists began to study the long-term, large-scale effects of SST operation on the Earth’s atmosphere, atmospheric change emerged as a meaningful global environmental issue for the first time. The truly global scale of the atmosphere and the systems-based methodology scientists used to study it distinguished atmospheric change

136 New York Times, “Johnson Gets Report,” May 4, 1965.

67 from local and regional environmental problems like land use, air pollution, and water quality. Only scientists had the tools and expertise to identify threats to the atmospheric environment, and as gatekeepers to information about the global atmosphere, they were forced to take a more active role in American environmental politics.

They did so with some ambivalence. Many scientists shared American environmentalists’ concerns over the unintended consequences of technological development, but as a group they also relied on high-tech tools and government institutions to advance their research and to do their jobs. They disagreed with what they perceived as environmentalists’ more general reaction against technology. In fact, both scientists and environmentalists struggled to reconcile their fascination with and dependence on cutting-edge modern technologies, and their differences had as much to do with environmentalists’ distrust of government-sponsored scientific projects and institutions as it did with their objections to technology itself.

As the debate over the SST unfolded, atmospheric scientists developed a particular form of environmental advocacy that reflected their professional values. Their authority as experts rested on their political neutrality, and they generally shied away from specific policy proscriptions or public appeals that might undermine their commitment to “good science.” Their efforts to promote the interests of their disciplines and institutions within the wider scientific community shaped their approach to environmental politics. Environmentalists generally sought to mobilize middle-class citizens in grassroots campaigns of lobbying, legislation, and litigation on specific, often local environmental issues. Atmospheric scientists, by contrast, favored a national and international scale bureaucratic approach. Cutting-edge knowledge disseminated to

68 capable administrators, scientists believed, would lead to better scientific research and, eventually, to sound environmental policy. They sought to gain influence among the scientific and governmental elites that helped determine national and international scientific priorities and research agendas in order to expand research into the global atmosphere. Though they framed atmospheric change as part of a global environmental crisis and genuinely worried about threats to the atmospheric environment, atmospheric scientists focused their efforts primarily on influencing the internal politics of science itself.

SST: Some Economic and Political Background

The Johnson administration did its best to keep debates about the economic viability and political desirability of the SST in-house, and until 1967, it largely succeeded. The FAA and the Kennedy-created Presidential Action Committee on the

SST discussed the project at length, but the Administration used concerns over security to circumvent public scrutiny of the project.137 The plane’s status as a Kennedy-era aerospace project also helped insulate the SST from both Congressional and media opposition. In the post-Sputnik aerospace technology boom of the early 1960s, few

Senators had incentive to stand in the way of a program its promoters claimed would create American jobs, bolster the American economy, and ensure American hegemony in aircraft manufacture in the face of international competition. Washington State, the home of SST’s primary design firm Boeing and a major hub of aircraft manufacturing in the

United States, stood to benefit significantly from the project. The state’s powerful

137 Joshua Rosenbloom, “The Politics of the American SST Programme: Origin, Opposition, and Termination,” Social Studies of Science, Vol. 11, No. 4 (Nov. 1981), 403-423.

69 Democratic senators, Warren Magnuson and Henry “Scoop” Jackson, lent their strong support. Meanwhile, the Anglo-French Concorde, as well as the Russian TU-144, seemed likely to hit production by the early 1970s. Proponents of the SST reiterated time and again that Americans were either going to be building supersonic jets in the decades to come or they were going to have to buy them.138

Despite this early public support, the SST’s economic liabilities quickly earned the program critics. As Swedish aviation consultant Bo Lundberg wrote in a 1965

Bulletin of the Atomic Scientists article (and later testified to Congress), while the SST might create a few jobs for highly-skilled workers already in high demand, the project held little real promise of bolstering the U.S. economy.139 It might not even break even.

Lundberg predicted that the Concorde, the model for the American SST, would at the projected cost $46 million per plane and 3-4 times the fuel consumption per seat mile of a standard subsonic jet, serve an “appallingly small” market. In the end, he presciently contended, the Concorde would lose its British and French sponsors money.140 The

American SST might fare even worse. Aware of Lundberg’s criticisms, Defense

Secretary Robert McNamara had his own team of analysts investigate the plane’s long- term business plan, and was similarly unimpressed with the results. He quietly sought to

138 In the days before Kennedy’s June 5, 1963 announcement, Pan American had made its order of six Ango-French Concordes public, spawning one of many waves of this type of rhetoric. A presidential fact sheet on the SST, put together by Alexander Butterfield in September of 1969, declared that it was “only a matter of a few years before SSTs will be flying inter-oceanic and trans-polar routes. If the President had not moved ahead we would have found it necessary later on to spend hundreds of millions of dollars abroad buying SSTs made by others. Now we can build them ourselves and sell them abroad…and thereby strengthen our balance of payments position.” Butterfield to John C. Whitaker, Nixon Presidential Materials Project, White House Central Files, Staff Member Office Files, John. C. Whitaker Papers; Rosenbloom, 405. 139 B.K.O. Lundberg, “Supersonic Adventure,” Bulletin of the Atomic Scientists, Vol. 21 (February, 1965), 29-33; U.S. Congress, Joint Economic Committee, The Supersonic Transport, Hearings, 92nd Congress, 2nd Session, December 27 and 28, 1972. 140 U.S. Congress, The Supersonic Transport, 63.

70 kill the economically dubious program behind an enthusiastic President Johnson’s back.141 As Erik Conway demonstrates in his comprehensive analysis of the

“technopolitics” of supersonic transport, little demand for the new planes existed outside of the airline manufacturing industry, and by 1968, even Boeing, the plane’s chief

American designer, saw the project more as a technological test venture than as a commercially viable craft.142

The SST, like the Concorde, was a “political aeroplane,” and carried with it a set of assumptions about the role of technological development in foreign policy and domestic economic stability. At least until 1967, these trumped Lundberg’s and

McNamara’s hard-nosed economic analyses.143 With the political momentum of the

Space Race and the attendant aerospace technology boom behind it, the American SST might have become a reality had not a problem with Boeing’s wing design stalled the

141 Kenneth Owen provides a detailed account of McNamara’s resistance to the SST in his investigating of the international politics of Supersonic Transports, Concorde and the Americans: International Politics of the Supersonic Transport (Washington, D.C.: Smithsonian Institution Press, 1997). Also see Rosenbloom, 406. 142 Conway’s use of internal Boeing documents and trade paper reports, in combination with Congressional hearings and other more traditional public records, makes his High Speed Dreams: NASA and the Technopolitics of Supersonic Transportation, 1945-1999 (Baltimore: Johns Hopkins University Press, 2005) a fascinating and authoritative account of the SST controversy in America. His work generally supports Mel Horwitch’s argument that environmental concerns provided the key locus of public protest against the SST, but augments that analysis with a clearer picture of the internal conflicts over SST within the business community and the Kennedy, Johnson, and Nixon administrations. Mel Horwitch, Clipped Wings: The American SST Conflict (Cambridge: MIT Press, 1982). 143 Andrew Wilson, Aviation Correspondent for the London Observer, outlined the international political context of the Concorde’s development before Congress in 1972, as well as in a 1973 paperback, The Concorde Fiasco (Middlesex: Penguin Books, 1973). Begun in 1962, the Concorde project was initially expected to cost about $400 million, resulting in 400 operational airplanes by 1990. In reality, the joint venture was in part a bid by the British to garner French support for their entrance into the European Economic Community. The British bid was eventually vetoed…by the French. The Anglo-French Concorde contract contained no escape clause, leaving the British to support the anemic craft until it was finally retired in 2003. The Anglo-French sales strategy, Wilson recalled, was “nakedly one of blackmail,” whereby the Concorde’s attraction of first class customers threatened to destroy traditional first-class operations for other airlines. In order to compete, airlines would have to buy a Concorde themselves. Curiously and comically, the cover of Wilson’s paperback features an enormous plush caricature of a Concorde wearing a businessman’s suit and a precarious bowler hat, with its big eyes lidded and a bent cigarette butt hanging from its mouth. U.S. Congress, The Supersonic Transport, 89.

71 project long enough to allow doubts about the plane’s economic viability—and about these Kennedy-era pro-technology assumptions—to catch up.

Opposition to the SST mounted in 1967 during the project’s delay, and the fight against the plane began to represent a larger challenge to the mindset that had made the

SST a possibility in 1963.144 By the time the project was defeated in 1971, Democratic senators cited the SST as a “symbolic issue in the struggle over new priorities and directions for the nation.”145 Much of the impetus for this “struggle for new directions” came from a deep skepticism about the integrity of government and the role of an overly powerful technology sector tied directly to America’s controversial military engagements. Opposition that began as an economic argument against irresponsible aerospace spending, however, came to fruition in the form of a growing concern for the quality of America’s natural and human environments.

Sonic Booms, Noise Pollution, and the Defeat of the SST

In Beauty, Health, and Permanence, Sam Hays has argued that concerns about threats to American’s everyday quality of life drove the post-war American environmental movement. For many environmentalists, the Supersonic Transport represented just such a threat. The SST was louder, less fuel efficient, and more costly than any other aircraft of its day, but for citizens’ groups and environmentalists who opposed the SST, the plane’s foremost liability was the deafening shock wave it left in its wake. Traveling at speeds exceeding Mach 1 (the speed of sound, about 340 m/s, or about 760 mph at sea level) left a rumbling blanket of disruption over the landscape

144 Rosenbloom, “Politics of the American SST,” 411. 145 John W. Finney, “Miscalculations by White House and Labor Helped Defeat Supersonic Transport,” New York Times, March 25, 1971.

72 below, a sharp and crackling clap that shook windows and in some cases even damaged buildings.146 The SST promised to visit this disruption upon thousands of communities across the nation on a daily basis. The boom provided the environmental movement with a ready-made “quality of life” argument against the program.

In 1967, popular opposition to the SST coalesced around two Harvard professors, biologist John Edsall and physicist William Shurcliff. Influenced by Lundberg, who by

1966 had expanded his criticism of the SST to include the sonic boom issue, Shurcliff founded the Citizens League Against the Sonic Boom in March of 1967, with himself as director and Edsall as his deputy.147 Shurcliff lobbied Democratic members of Congress to put pressure on the FAA and NASA to reassess the SST in light of a series of tests conducted over Oklahoma City that demonstrated just how disruptive daily sonic booms could be.148 He complained that the tests had not been adequately considered in NASA’s pro-SST report on the sonic boom.149

While Shurcliff lobbied Congress to reevaluate the Oklahoma City tests, he also began to pitch the sonic boom problem to American environmentalists—particularly the

Sierra Club and the Wilderness Society. In order to sell the cause to organizations like the Sierra Club and its “Group of Ten” environmental associates, Shurcliff noted that sonic booms would disturb wildlife and the tranquility of wild lands. Secretary of the

146 At 65,000 feet, where the temperature is around -70°F, the speed of sound actually drops to about 660 mph. When sound is emitted from a stationary object, it creates an oscillating differential of air pressure that travels in waves and creates the phenomenon of sound in our ears. When an object, like, for example, a Supersonic Aircraft, travels faster than the speed at which these waves propagate, the result is a compression of air around the front edges of the object. That compression emits a conical shock wave which in turn causes a sudden change in air pressure on the ground that our ears perceive as a sound much like thunder, a sonic boom. Horwitch, Clipped Wings,74. 147 Horwitch, Clipped Wings, 222; Conway, High Speed Dreams, 126-127; Owen, Concorde and the Americans, 114-115. 148 For more on the Oklahoma City tests, see Conway, High Speed Dreams, 121-125; Horwitch, Clipped Wings, 75-79. 149 Horwitch, Clipped Wings, 222.

73 Interior Stewart Udall echoed Shurcliff’s concerns in an article on the front page of the

Washington Post, where he publicly “decried the boom’s impact on wildlife, national parks, and adobe structures built in the Southwest by Indians.”150 Though marginal compared to more prominent environmental affronts like logging and mining, the threat that sonic booms presented to both Indian heritage and the American wilderness experience resonated among many members of America’s environmental groups.

The Citizens’ League’s fight against the SST played out within the familiar patterns of legislative lobbying, litigation, and environmental monitoring that have come to characterize modern environmental politics. Like the bulk of the environmental activism of the period, opposition to the sonic boom began with a grassroots organization representing a large constituency of concerned citizens and flowered into a national campaign to protect Americans’ overall quality of life. The Sierra Club, the Wilderness

Society, and later Friends of the Earth, Environment!, NRDC, and other major environmental groups all moved to oppose the SST. Their campaigns achieved notable success.

Environmentalists quickly expanded their criticism of the SST to include the related issue of airport noise pollution. In addition to producing sonic booms while in flight, the SST engines roared at decibels well above those dictated by new airport noise regulations on the ground. Citizens groups had already convinced the FAA to begin regulating conventional aircraft noise at America’s airports, and the SST would negate these gains. Like the sonic boom, airport noise threatened American’s day-to-day quality of life.

150 Don Dwiggings, The SST: Here it Comes Ready or Not, New York: Doubleday, 1968, pg. 66. David Hoffman, “Would the SST Peril U.S. Tranquility?” Washington Post, December 20, 1967.

74 Ironically, Boeing, the Seattle-based firm commissioned to design and build the

SST, inadvertently helped to bring the issue of airport noise pollution to the fore when it introduced the 747 “jumbo” in 1967. The introduction of jet engines to regional airports had spawned concerns over noise pollution in the mid-1960s, but the perceived possibility of 747s on every tarmac pushed the limits of what citizens living near airports were willing to tolerate.151 As early as 1965—before the “jumbo’s” release—Johnson’s

Science Advisor Donald Hornig, at the behest of members of Congress, put together a noise abatement advisory committee. After a brief turf war between the FAA and the Air

Transportation Association over regulatory tactics and jurisdiction, President Johnson signed a noise abatement bill into law in the summer of 1968, setting specific standards for the jumbo that the FAA had developed based on recommendations from, of all places,

Boeing. Even by Boeing’s own accounting, however, the SST stood no chance of meeting the jumbo’s noise standards, at least not with the General Electric GE4 engine that the craft had been designed around. Improvements in the efficiency of subsonic jet engines—particularly the new high-bypass turbofan engines in the Douglas DC-8 and the

Boeing 747—went hand in hand with noise reductions, but the restricted airflow of the

GE4 meant that this relationship between efficiency and noise was reversed for the supersonic. The better the SST ran, the louder it roared.152 Fittingly, Johnson’s noise bill also limited sonic booms, a tacit admission of at least temporary defeat on the SST by a beleaguered President.153

151 For more on the Boeing 747, see Conway, High Speed Dreams, 134-5; Horwitch, Clipped Wings, 158- 161, 174. 152 Conway, High Speed Dreams, 9, 138. 153 New York Times, “Johnson Signs Bill to Curb Jet Plane Noise and Booms,” July 25, 1968; Conway, High Speed Dreams, 137.

75 By all rights Johnson’s 1968 noise bill should have finally killed the SST. The sonic boom issue had already raised the public’s ire toward the project, and Johnson’s increasingly secretive management of the FAA regarding supersonics prompted journalists to equate the SST with his questionable prosecution of the controversial and unpopular Vietnam conflict. As concerns over inflation and war spending forced governmental agencies to tighten belts, and as new political pressures from environmental and citizens’ groups began to change Congressional thinking about the political viability of the aerospace pork barrel, a supposedly new set of national priorities should have had little room for civilian supersonics.154

The SST, however, had considerable political momentum. More than a decade of

Cold War technological jingoism, hundreds of millions of dollars, and immeasurable political capital had been sunk into the project. NASA, the Department of

Transportation, and the FAA had no interest in watching the SST die. And neither, it turned out, did the new President of the United States, Richard Nixon.

Nixon sought to push forward with the SST for two main reasons. First, despite the plane’s economic liabilities, Nixon’s advisors feared that without an SST, the U.S. would “take a back seat in aviation history for the next decade.”155 The Johnson

Administration had invested considerable sums of money into the project, and the Nixon

Administration hoped to capitalize both economically and politically on that investment.

Nixon and his staff recognized the program’s short-term economic limitations, but they expected that a functioning SST would create its own lucrative market by adding a new

154 See Linda R. Cohen and Roger G. Noll, The Technology Pork Barrel (Washington: The Brookings Institution, 1991). 155 Alexander P. Butterfield to John C. Whitaker, memorandum, September 23, 1969, John C. Whitaker Papers.

76 dimension to commercial aviation. Nixon’s Deputy Assistant, Paul Butterfield, predicted that once supersonics began flying regular schedules between the U.S. and Europe, their routes would expand to Asia and South America. The market for SSTs would expand in turn, giving the nations that manufactured the planes an advantage in their “balance of payments” position.156 Supported by this economic rationale, the Nixon Administration reiterated Johnson’s old argument that the U.S. would either have to build supersonics or buy them.

Nixon’s economic rationale was tenuous, but he also had a political incentive to continue the SST program. The Administration worried that canceling the SST would draw the ire of American labor unions, with whom Nixon already had a prickly relationship. Nixon’s staff was in the process of trying to curb what the President considered “inflationary” wage increases proposed by manufacturing unions, and Nixon had no interest in further alienating workers by cutting high-profile American aerospace jobs during his first few years in office.157 That those jobs were mostly in Washington

State—again, the home of two powerful Democratic Senators, made the prospect of terminating the SST even less palatable.

Nixon understood that the political climate required a certain budgetary frugality on the SST for fiscal year 1970, but he took steps to keep the plane’s design and construction contractors “on standby.” He instructed his eager Secretary of

Transportation Harold Volpe to continue discretionary payments of $11 million each month to Boeing and General Electric to maintain the project’s momentum.158 While he

156 Ibid. 157 Memorandum for the President, April 24, 1969, John C. Whitaker Papers. 158 Volpe may have been a little over-eager in his meeting with the new President. As Whitaker recounted in a memorandum to the President summarizing the meeting, Volpe drew an analogy between Nixon’s SST

77 was unwilling to grant Volpe’s $200 million FY 1970 request for SST development, he didn’t blink at the $1.5 billion in long-term government funds that Volpe projected for supersonic development through the first few prototypes.159 Over the next sixth months,

Nixon’s staff developed a plan to reintroduce the government-financed SST project, announcing in September the move from Phase II of the project—the design phase—to

Phase III, the actual “cutting of metal” on prototype models.160

Unlike the Johnson administration, Nixon and his staff actively sought to preempt criticism of the project. The Administration released “informational fact sheets” on the benefits of the SST to American consumers, to the nation’s balance of payments, and to the economy in general. In addition, having created and hand-picked the Council on

Environmental Quality in August of 1969, Nixon sought to quiet his critics’ fears about the environmental effects of SST by incorporating that body’s recommendations into the plane’s development.161

The administration’s strategy backfired. Acting head of the CEQ Russell Train, whom Nixon had tapped as much for his ties to the Republican Party as for his knowledge of and concern over the natural environment, apparently took his job more seriously than the President expected him to. In May of 1970, at a series of hearings hosted by Senator William Proxmire of Wisconsin to capitalize on the success of Earth

Day, Train publicly summarized the myriad potential environmental problems of the

SST. He vowed that the administration would only put the craft into production if it

support and Columbus’s spirit of discovery, “an analogy that seemed to embarrass” Nixon. Memorandum for the President, April 24, 1969, John C. Whitaker Papers. 159 Ibid. 160 Memorandum for John C. Whitaker, September 23, 1969, John C. Whitaker Papers. 161 Memorandum for Environmental Quality File, August 9, 1969, John C. Whitaker Papers.

78 could resolve these heretofore-irresolvable issues.162 Compounding the damage, the respected physicist Richard Garwin, who had chaired Nixon’s ad hoc Office of Science and Technology Committee on the SST in 1969, also attacked the program, both for its environmental impacts and for its economic inviability. His testimony echoed criticisms from his own OST report—a report that Nixon had conveniently squelched in 1969.163

Train committed to an environmentally acceptable SST, but Garwin concluded that such a craft could not be built. By the time Nixon’s 1971 funding request hit the Senate floor in November of 1970, most of the nation’s renowned scientists and economists, the CEQ,

Nixon’s own Science Advisor Lee Dubridge, perhaps a majority of the American public, and even some inside Boeing, opposed the program.164 The Senate killed SST’s funding by a narrow vote on March 24, 1971.165

SST and the Atmosphere: A New Type of Environmental Issue

Sonic booms, noise pollution, and economics killed the SST, but the debate over the plane also introduced important new players and a new type of global environmental issue to American environmental politics. The players were scientists, and the issue was atmospheric change. Between 1968 and 1971, an increasingly vocal group of atmospheric scientists raised a new set of concerns about the effects of high-elevation

162 U.S Congress, Joint Economic Committee, Subcommittee on Economy in Government, Hearing, Economic Analysis and the Efficiency of Government. Part 4: Supersonic Transport Development, 91st Congress, 2nd Session, May 7, 11, and 12, 1970 (Washington, D.C.: U.S. Government Printing Office,1970): 904-920; also cited in Conway, High Speed Dreams, 142. 163 Garwin had a long and storied career, both as a scientist and as a science advisor. Most notably, Edward Teller gave him credit, at 23, for the 1952 drawings that led to the creation of the hydrogen bomb. Earl Lane, “Physicist Richard Garwin: A Life in Labs and the Halls of Power,” American Association for the Advancement of Science, January 17, 2006; http://www.aaas.org/news/releases/2006/0117garwin.shtml. 164 Both Milton Friedman and John Kenneth Galbraith, renowned economists who rarely agreed on anything, thought the SST a bad idea. Conway, High Speed Dreams, 144. 165 Finney, “Miscalculations by White House.”

79 supersonic flight on the Earth’s atmosphere—effects that potentially included ozone depletion, increased solar radiation, and climate change. In fact, Train’s testimony proved damaging not only because he committed the administration to tackling the unsolvable technical problems of booms and noise, but also because he expressed concern over recent studies of the SST’s effects on the upper atmosphere—effects that previous SST committees hadn’t yet considered.166 Scientists at government-funded institutions like NASA and the National Center for Atmospheric Research (NCAR) conducted research into the SST’s atmospheric impact in part to fulfill institutional missions to conduct more socially relevant research. But by framing their concerns about the SST’s impact on the atmosphere in environmental terms, these scientists also helped to bring atmospheric change into the realm of American environmental politics in the

1970s.

Atmospheric research had been steadily growing in strength and appeal over the course of the 1960s, and the SST issue provided a much-needed application of the discipline’s basic research. Like other American scientists, atmospheric scientists faced a number of challenges in the new decade. A waning of the aerospace era coincided with an overall increase in new Ph.D.s (especially in physics), and many scientists expressed concerns over the role of American science in the Vietnam War. In 1970s, the discipline’s institutional leaders were looking for new directions.167

166 U.S. Congress, Economic Analysis, May 7, 1970, 1000. 167 For more on the institutional development of atmospheric science in the 1960s, see chapter 1. For the physics community at large, see Daniel J. Kevles, The Physicists: The History of a Scientific Community in Modern America (Cambridge: Harvard University Press, 1995), and for the Ph.D. glut specifically, see David Kaiser, American Physicists and the Cold War Bubble (Chicago: University of Chicago Press, forthcoming).

80 At NCAR, measuring the SST’s effects on atmospheric composition directly addressed both an institutional mission to study the myriad aspects of the Earth’s atmosphere and growing pressure from the National Science Foundation to demonstrate the practical application of atmospheric and climatic research. In 1970, Nixon created the National Oceanic and Atmospheric Administration, placing the new agency’s headquarters practically down the street from NCAR in Boulder. NCAR was technically a non-governmental institution, but it nevertheless relied heavily on federal funding. The institution’s leaders looked upon NOAA warily. Herb Hollman of the Department of

Commerce (NOAA’s parent agency) fueled their fears. He had no direct authority over the NSF—also a federal agency—but he nevertheless threatened to dislodge NCAR from its NSF moorings and roll it into NOAA unless NCAR made an effort to “pay more attention to the public policy side” of atmospheric research.168 For NCAR’s concerned leaders, SST research provided one way to demonstrate the institution’s larger social value.

Scientifically, NCAR was especially primed for studying the SST. Over the course of the 1960s under directors Walt Roberts and John Firor, the institution’s leaders actively sought to move the atmospheric sciences beyond traditional meteorology. They hoped to integrate chemical, physical, and numerical expertise in an effort to create both theoretical and numerical models of atmospheric circulation at a number of geographical and chronological scales.169 Both Roberts and Firor believed that integrated atmospheric

168 Interview with John Firor, June 26, 1990, by Earl Droessler, UCAR/NCAR Oral History Project. 169 Together, they helped promote the Global Atmospheric Research Project (GARP), a joint project of the World Meteorological Organization and the International Council of Scientific Unions designed chiefly to improve short-term weather predictions through a broad network of data collection, modeling, and experimentation. The project also incorporated research into the long term physical, chemical, and biological components of climate, and included a number of NCAR scientists. For more on GARP, see

81 research at NCAR should serve the “national interest,” and NCAR’s scientists and managers continued to frame their projects in terms of socially relevant issues like air pollution or weather modification.170 As early as 1965, the director of NCAR’s

Laboratory of Atmospheric Sciences, William Kellogg, had taken an interest in monitoring atmospheric ozone using satellites, discussing possible partnerships with

NASA or Bell Labs to build the necessary instruments.171

Kellogg’s collaborations with other scientific institutions also reflected a growing interest in the atmosphere among a broader scientific community. At NASA, the final

Apollo missions represented the end of the seemingly limitless resources of Kennedy-era space exploration, and with the shuttle program nearly a decade away, the agency began a temporary turn earthward, using its orbiting and earthbound resources to begin monitoring the Earth’s atmosphere in what would become the Upper Atmosphere

Research Program.172 Theoretical concerns about SST’s effects on stratospheric ozone also cropped up among Kellogg’s colleagues in the National Academy of Sciences in

1965, and in computer models at the Weather Bureau’s Geophysical Fluid Dynamic chapter 1. See also Erik Conway, A History of Atmospheric Science at NASA, 1958-2004 (Baltimore, MD: Johns Hopkins University Press, forthcoming), 94-144; Spencer Weart, The Discovery of Global Warming (Cambridge: Harvard University Press, 2003), 99-100. 170 Would-be Washington Governor Dixie Lee Ray, representing the NSF in medical and biological sciences at the time, questioned the idea of defining the “national interest” in an unexpected visit to NCAR in 1963. From internal memos about the visit, the energetic Stanford-trained marine biologist seems somewhat comically to have inspired a certain amount of fear among the fledgling NCAR’s leadership. Memorandum by Al Morris, “Visit of Dr. Dixie Lee Ray of NSF,” January 22, 1963, Philip D. Thompson Papers, UCAR/NCAR Archives, Boulder, CO; Interview with John Firor, June 26, 1990. 171 Will Kellogg to Walt Roberts, memorandum, “Possibility of a Satellite Experiment to Determine Ozone by Dave,” Philip D. Thompson Papers. 172 Stephen Schneider discusses these concurrent shifts from the perspective of an advanced graduate student working between Columbia University and NASA’s Goddard Institute for Space Studies in New York, just before being courted by NCAR in Boulder. Interview with Stephen Schneider by Bob Chervin, January 10-13, 2002, AMS/UCAR Tape Recorded Interview Project. NASA’s Earthward shift, as Conway points out in Atmospheric Science at NASA (160-161), did come as a response to the end of the Apollo missions, but it may not have been quite the departure that it first appears. Much of the impetus for the Upper Atmosphere Research Project, he notes, arose from the need to file a meaningful environmental impact statement for the agency’s new shuttle project. NASA, after all, was the government’s lead agency on all things aerospace, and its administrators saw the SST as part of its institutional responsibilities.

82 Laboratory in Princeton around the same time.173 For Kellogg and other leaders of atmospheric science, the SST-ozone problem could not have been more appealing.

Kellogg began to study the SST in earnest in 1968, just as the first wave of environmental concern over the project began to plateau. Under pressure from Hollman and the NSF to make his basic research more relevant and applicable, Kellogg worked with scientists from institutions around the world to discover if water vapor, carbon dioxide, and/or nitrogen oxides released by a fleet of 500 Boeing SSTs could deplete the ozone layer or cause climate change, as some atmospheric scientists, like James

McDonald of the University of Arizona, feared.174 In 1968, McDonald began to explain how, at least theoretically, the SST might increase the amount of water vapor and other gases in the upper atmosphere significantly enough to alter the climate and deplete stratospheric ozone.175 A decrease in ozone, he warned, could increase the amount of radiation that made its way to Earth and thereby significantly increase the risk of skin cancer. While most scientists agreed that a reduction in ozone would lead to an increase in cases of skin cancer, many, including Kellogg, were skeptical of McDonald’s basic claims about ozone depletion.176

173 National Research Council Committee on Atmospheric Sciences, Atmospheric ozone studies; an outline for an international observation program, a report by the Panel on Ozone to the Committee on Atmospheric Sciences National Academy of Sciences, National Research Council, November, 1965 (Washington, D.C.: National Academy of Sciences-National Research Council, 1966). 174 Interview with John Firor, June 26, 1990. 175 Conway, Atmospheric Science at NASA, 201. 176 McDonald is an interesting character is the history of science. Not only was he a very perceptive atmospheric scientist, who, despite being challenged, articulately expressed valid concerns about the effects of SST on the environment, he was also a strong and public proponent of UFO research, and promoted the controversial “extra terrestrial hypothesis of UFOs,” which significantly sullied his reputation among many in the scientific community. Many scientists, however, found his call for scientists to responsibly study UFOs compelling. He famously gave an address entitled “Science in Default” on the subject to the AAAS. Eventually, however, he became isolated within the scientific community and increasingly unhappy. His wife decided to divorce him in 1971, and he attempted suicide, succeeding on his second attempt in June of 1971. For McDonald’s testimony, see U.S. Congress, House Committee on Government Operations, Hearing, FAA Certification of the SST Concorde, 94th Congress, 1st Session, July 24, November 13, 14,

83 In 1970, after an MIT-sponsored conference on “Man’s Impact on the Global

Environment,” Kellogg and his colleagues presented their SST research in a chapter of the Report on the Study of Critical Environmental Problems, or SCEP.177 Organized primarily by Carroll L. Wilson of MIT’s Sloan School of Management (and of MIT’s

Program for the Social Application of Technology), SCEP established the parameters for a new kind of top-to-bottom integrated scientific assessment of the human environment, and recommended a scientific agenda for the upcoming United Nations Conference on the Human Environment in 1972.178 Despite the paucity of vertical mixing in the upper atmosphere, allowing the gases of SST’s exhaust to stay airborne for between one and three years, the SCEP group concluded that “no problems should arise from the introduction of carbon dioxide and that the reduction of ozone due to interaction with water vapor or other exhaust gases should be insignificant.”179 But the SCEP scientists also recognized that the study had relied on General Electric’s data for its own engine’s emissions, and their report contained many uncertainties. They recommended “that uncertainties about SST contamination and its effects be resolved before large-scale operation of SSTs begins,” and outlined a brief but expansive research plan for obtaining better data on the subject. 180

December 9, 12, 1975, February. 24, 1976 (Washington, D.C.: U.S. Government Printing Office, 1975): 93. For a sympathetic biography of McDonald, see Ann Druffel, Firestorm: Dr. James E. McDonald's Fight for UFO Science (Columbus, NC: Wild Flower Press, 2003). 177 Man’s Impact on the Global Environment: Assessment and Recommendations for Action, Report on the Study of Critical Environmental Problems (Cambridge: MIT Press, 1970), hereafter referred to as “SCEP”; Stuart Auerbach, “Scientists Fear Climate Change by SST Pollution,” Los Angeles Times, August 2, 1970. It should also be noted that Auerbach, though reporting for the LA Times here, wrote predominantly for the Washington Post. 178 For more on SCEP and the U.N. Conference on the Human Environment, see chapter 3. 179 SCEP, 16. 180 SCEP, 17. George D. Robinson, then of the Center for the Environment and Man, Inc (a nearly all- encompassing name, it appears), remembers SCEP as a “visionary report man’s impact on the environment,” but one without the poetry in it. “It was visionary in the sense you are you ever going to get anyone to pay for all this, but it wasn’t visionary in the sense that it didn’t tell you how much it would cost.

84 In August of 1970, before the official publication of the SCEP report, the overtly anti-SST New York Times and LA Times ran articles on the SCEP conference proceedings, playing up the recommendation that the project be delayed. The story made the front page of both papers, with the LA Times declaring “Scientists Fear Climate

Change by SST Pollution,” citing concerns about CO2 and other gases trapped in the stratosphere and quoting Kellogg specifically. “When you change something on a global basis,” Kellogg told the press, “you had better watch out.”181 Kellogg’s caution, however, was only half of the story; he and his SCEP co-authors certainly had concerns about both ozone depletion and climate change—and Kellogg would express deep concerns about anthropogenic climate change and environmental degradation throughout his life—but based on his research, he and his colleagues saw no direct evidence that the SST would cause either of these phenomena. In early March of 1971, as the Senate debated SST’s future, Kellogg testified before Congress that he had so far found no convincing evidence that the SST would cause significant ozone depletion or climate change.182 SST supporter

S. Fred Singer, then chairman of the Department of Transportation’s SST Environmental

Advisory Committee (and later a vocal denier of global warming), testified alongside

Kellogg, and concurred with his findings. “In short,” wrote Kellogg in his prepared statement, “I have found no environmental basis for delaying the Government’s SST

It does tell you how much it would cost.” Interview of William Kellogg by Ed Wolff and Nancy Gauss, February 10, 1980, NCAR.UCAR Oral History Project; Interview of George D. Robinson by Earl Droessler, June 27-28, 1994, AMS/UCAR Tape Recorded Interview Project. 181 Bayard Webster of The New York Times gave a less alarmist report, but still emphasized the delay rather than the overall conclusion about the SST. Webster, “Scientists Ask SST Delay Pending Study of Pollution,” New York Times, August 2, 1970. 182 U.S. Congress, House Committee on Appropriations, Hearing, Civil Supersonic Aircraft Development (SST), 92nd Congress, 1st Session, March 1-4,1971 (Washington, D.C.: U.S. Government Printing Office, 1971): 512-541; Christopher Lyndon, “Experts Assure House SST Would Not Be Harmful,” New York Times, March 4, 1971.

85 program…”183 But Kellogg also highlighted the vast uncertainties surrounding the SST’s effects on the atmosphere, and called for further research on the subject. If his testimony gave the forthcoming SST prototype a tentative green light vis-a-vis the atmosphere,

Kellogg’s cautionary and equivocal SCEP report—along with his comments to the press—hardly provided the plane’s backers with a ringing environmental endorsement.184

Harold Johnston, a University of California, Berkeley chemist, worried that

Kellogg and Singer underestimated the SST’s impact. Kellogg and Singer agreed that

SSTs might cause some ozone depletion and could potentially lower the temperature of the stratosphere, but they concluded that these changes would be insignificant compared to natural atmospheric variations. Johnston disagreed. He predicted a much greater reduction of ozone due to supersonic flights than the SCEP study anticipated, if not because of increased water vapor, because of the chemical reactions between various oxides of nitrogen and the extra oxygen atom of ozone molecules. At a Department of

Commerce panel on the SST held in March of 1971 in Boulder, Johnston objected to

185 SCEP’s conservative analysis of the NOx problem. An expert on ground-level ozone chemistry, Johnston claimed that trace gases could significantly reduce the amount of ozone in the stratosphere—enough, he worried, to increase significantly the risk of skin cancer around the world, as McDonald had feared.186 Though few scientists at the

Boulder conference made the jump to Johnston’s extreme case of up to a 50% reduction in ozone, Johnston’s concerns raised eyebrows. When he returned to Berkeley, he began

183 U.S. Congress, Civil Supersonic Aircraft Development, 524. 184 Kellogg appears to have recognized the myriad non-environmental drawbacks of the SST, and he “urge[d] that the SST issue be decided on grounds that are more significant than the effect on the environment.” U.S. Congress, Civil Supersonic Aircraft Development, 526. 185 U.S. Congress, FAA Certification of the SST Concorde, 93. 186 Ibid.; Conway, High Speed Dreams, 164.

86 writing up his research in a controversial paper that would appear in Science in August of

1971.187

Johnston’s results, leaked to New York Times science writer Walter Sullivan in late May of 1971, appeared only after Congress had voted to kill the SST, but his work nevertheless found an audience.188 In Europe, the Anglo-French Concorde was entering the construction phase, and further debates about Supersonics and their role in the

American air travel market lurked on the horizon. Now the Earth’s atmosphere took center stage. Working with NASA, the Department of Transportation created a four-year atmospheric monitoring program designed specifically (and belatedly) to study the atmospheric and climatic effects of the SST, the Climate Impacts Assessment Program, or CIAP.189 Vindicating both McDonald and Johnston, CIAP found that a fleet of 500

Boeing 2707 SSTs would in fact decrease ozone globally by about 12%, leading to an increase in skin cancer cases of approximately 24% in the United States, representing about 120,000 new cases per year, or 2.4 million cases over the 20 year lifetime of one of the planes. For the Concorde, which flies lower than the Boeing 2707 would have, the numbers were significantly lower, but the effect was essentially the same. The SST’s complex effects on climate still concerned some scientists as well. Amidst controversy over the certainty of their results, CIAP’s participants recommended further limitations

187 The controversial peer-review process is chronicled in Lydia Dotto and Harold Shriff, The Ozone War (Garden City, NY: Doubleday, 1978), 45, and in Conway, High Speed Dreams, 164; Harold Johnston, “Reduction of Stratospheric Ozone by Nitrogen Oxide Catalysts from Supersonic Transport Exhaust,” Science, Vol. 173, No. 3996 (Aug., 1971), pp. 517-22. 188 Walter Sullivan, “Sorry, but There’s Still More to Say on SST,” New York Times, May 30, 1971. 189 U.S. Congress, FAA Certification of the SST Concorde; See also Conway, Atmospheric Science at NASA, 205.

87 on Concorde’s U.S. flights, based in large part on its effects on the atmospheric environment.190

Research into the impact of the SST launched a fresh set of scientific and political conversations about a new type of “environmental” problem for the 1970s: the degradation of the Earth’s atmosphere. Both The New York Times and The LA Times categorized concerns over the SST’s impact on the atmosphere as fears about a new kind of pollution. They may have misrepresented the substance of Kellogg’s findings, but their characterization of climatic change and ozone depletion as new sorts of environmental issues accurately reflected the commonalities between scientists’ and environmentalists’ concerns.191 Kellogg’s “you better watch out” warning about the unintended consequences of modifying global systems echoed a sentiment common amongst environmentalists that threats to nature and nature’s systems should play a role in decisions about technological and economic development.192 The SCEP report explicitly underscored this commonality, framing its focus on global environmental problems as a complement to “local and regional problems…that many organization are deeply concerned with studying and ameliorating.”193

The scale of atmospheric change made it unique among mainstream environmental issues of the 1960s and ‘70s. Land use and waste management issues

190 Predictably, the three-volume report was buried beneath a comparatively optimistic executive summary, leading, also predictably, to further controversy on the SST. In the end, only 20 of the craft were commercially produced and put into service, largely at the continuing expense of the British government. I will discuss CIAP and the controversy over its findings in the following chapters. Ibid., 96-101. 191 Kellogg’s journals from the period reveal that he neither opposed nor fully supported the SST, but he certainly found no convincing scientific grounds on which to block the program. Indeed, his main objections, as discussed below, concerned the way science was being used to support opposing positions on the SST, to the detriment, in Kellogg’s mind, of science itself. If anything, he was sorry to see the program killed in the name of science when in fact the scientific evidence against the plane was highly uncertain. Auerbach, “Scientists Fear Climate Change”; Webster, “Scientists Ask SST Delay.” 192 Auerbach, “Scientists Fear Climate Change.” 193 SCEP, 5.

88 typically forced environmentalists to focus on specific partitions of terrestrial or aquatic space, as did efforts to protect wilderness areas and species habitat. Water pollution, though part of dynamic and theoretically global aquatic systems, generally stayed within the confines of watersheds and drainage patterns, and even common air pollution due to industry and agriculture—itself a form of atmospheric problem—was recognized as primarily a regional concern. An international environmental movement focused on resources and global environmental degradation grew up in tandem with American environmentalism during the 1960s, but many of the international movement’s primary concerns represented no more than a global summation of these key regionally-specific issues.194 Among environmental activists in the U.S. and abroad, NASA’s images of the

Earth from space helped shape an ideal of cooperative global environmental protection articulated in the “Only One Earth” theme chosen for the 1972 United Nations

Conference on the Human Environment.195 But when former Sierra Club Director and

Friends of the Earth founder David Brower famously exhorted environmentally- conscious Americans to “think globally, act locally,” it reflected a reality in which “the environment” that activists like Brower sought to protect was in fact an amalgamation of

“environments” discrete in space and time. The chemical constituents of the stratosphere, by contrast, actually circulated globally throughout the thin layer of gases encircling the

Earth. CO2, Ozone, and NOx, though measured in parts per million and parts per billion, resided semi-permanently as potential environmental offenders in a boundless global

194 This international movement will be discussed in some detail in the following chapter. “Declaration of the United Nations Conference on the Human Environment,” United Nations Environment Programme, http://www.unep.org/Documents.Multilingual/Default.asp?DocumentID=97&ArticleID=1503. 195 Stewart Brand, who will be discussed below, led a successful campaign in 1966 and 1967 to convince NASA to release the images. He thought, quite rightly, that they would provide a powerful symbol of Earth’s isolation that environmentalists could use to their advantage in the years to come. Andrew Kirk, Counterculture Green: The Whole Earth Catalog and American Environmentalism (Lawrence: University Press of Kansas, 2007).

89 space.196 Traveling at the speed of sound through the gaseous mixture of the atmosphere, an SST might alter the environment in its global totality, and the effects of that alteration might be felt anywhere…or everywhere.197

The uncertainties of large-scale, long-term climatic change further distinguished the atmosphere from other aspects of the global environment. The causal link between

SSTs, ozone reduction, and skin cancer, though controversial, presented environmentalists with an easily identifiable public health concern, albeit one difficult to connect to specific individuals or groups. The human and biological impacts of an overall change in the temperature of the stratosphere were more difficult to grapple with.

The impacts of climatic change on the biosphere might occur over decades—if not centuries. Changes in the atmosphere and climate transcended what the SCEP authors referred to as the “first-order” effects of other large-scale issues like population growth and pollution. Instead, the health of the stratosphere and the stability of the climate helped constitute an overall “status of the total global environment” that included a measure of potential environmental consequences at the third or fourth levels of causation.198 The Study of Man’s Impact on Climate (SMIC), a follow-up to SCEP, identified myriad ways in which humans inadvertently altered the climate at local, regional, and global levels, but the study only vaguely identified the reciprocal impacts of climatic change on humans, and these potential threats were laced with the caveats and probabilities of scientific uncertainty. In stark contrast to the sonic booms and airport

196 Some atmospheric constituents—especially particulate matter—were recognized to be distributed in zonal patterns by latitude that corresponded to the Earth’s prevailing winds. In the lower atmosphere, this zonal distribution is more pronounced. SMIC, 52. 197 Samuel P. Hays, Beauty, Health, and Permanence: Environmental Politics in the United States, 1955- 1985 (New York: Cambridge University Press, 1987), 8. 198 SCEP, 5.

90 noise issues that attracted environmentalists to the anti-SST camp in the 1960s, as a mainstream environmental issue, climatic change presented precious few direct, tangible threats to either human health or the natural world.

Even more so than SCEP, the SMIC report positioned humanity’s impact on the atmosphere as a mainstream environmental issue, both in content and in presentation. An offshoot of SCEP’s chapter on the “Climatic Effects of Man’s Activities”—written in large part by Kellogg—the 1971 SMIC report set out to “raise the level of informed public and scientific discussion and action on global and regional climate problems.”

Again, both Wilson and Kellogg helped to organize the meeting and served as authors.

SMIC tackled anthropogenic climate change not just from supersonics, but from automobiles, industrial pollution, and agricultural practices as well.199 After defining and describing a general theory of the Earth’s climate, the study moved systematically through humanity’s impact on climate from the ground up, discussing changes in the snow, ice, water, and dust at the Earth’s surface before moving on to clouds and trace gases in the troposphere and then to the all-important rarified air of the stratosphere. Like

SCEP, SMIC’s primary recommendations involved science policy (a $17.5 million global monitoring program). But rhetorically, SMIC also spoke to a larger group consisting of the “peoples and governments of the world who collectively bear the responsibility for assuring that man does not inadvertently destroy his environment in the process of meeting his many and varied needs.”200 With an iconic NASA image of the Earth from space gracing its cover and a Sanskrit prayer reading “Oh, Mother earth, ocean-girdled and mountain-breasted, pardon me for trampling on you” as the frontispiece, the

199 Inadvertent Climate Modification: Report of the Study of Man’s Impact on Climate (SMIC) (Cambridge: MIT Press, 1971), xv, xvii, hereafter referred to as “SMIC.” 200 SMIC, 3.

91 published study appealed to the precautionary ethos of the Earth Day teach-ins led by prominent environmental scientists and activists in April of the previous year.201 As a result of these efforts to study the impact of the SST on ozone and climate, scientific concerns about the atmosphere had by 1971 become part of a larger popular concern over the global environment. Atmospheric scientists themselves, however, hardly made typical environmentalists.

Scientists as Environmentalists?

The scale and complexity that distinguished atmospheric change from other environmental issues brought atmospheric scientists a prominent role in environmental politics.202 The long-term global effects of atmospheric change defied both the geography and the chronology of personal experience that provided the context for

American environmentalism. Scientific professionals dominated large-scale problems like population and global resource management, but in the 1960s and ‘70s these issues cropped up in everyday decisions about things like rising energy costs and family planning.203 Threats to the global atmosphere, by contrast, entered the public consciousness only through popularizations of scientific research. Identifying potential

201 The photograph on SMIC’s cover actually predates the “blue marble” image of Apollo 17, which has become the hallmark image of the earth. “Africa and the Mediterranean,” which graces the SMIC report, came from Apollo 11, the lunar landing mission launched on July 16, 1969. SMIC, cover and frontispiece. 202 Sam Hays argues that “environmental affairs” typically “stem from the circumstances of daily life, not from those shaped by technical specialization” (Beauty, Health, and Permanence, 8). In the years since the book’s publication, environmental expertise has risen in stature within environmental politics, muddling Hays’s formulation. In some instances, as in the case of climate change, the scale and complexity of environmental phenomena has predicated a stronger role for experts in defining environmental issues, but as Stephen Bocking demonstrates in Nature’s Experts: Science, Politics, and the Environment (New Brunswick: Rutgers University Press, 2006), even on a local scale, ecological expertise has, for better or worse (Bocking argues worse), become a primary tool for defining environmental values. Partly as a product of an increased focus on intangible problems like global warming and ozone depletion, environmentalists have increasingly begun to couch their arguments in scientific rather than moral or ethical terms, using the former to disguise the latter. 203 Hays, 9.

92 threats to this global space required both the expertise and the high-tech tools of

America’s atmospheric scientists. Without the international networks of data collection made possible by groups like the World Meteorological Organization and International

Council of Scientific Unions, in addition to the systems-based methodologies of the

SCEP and SMIC groups, environmentalists had no way of even recognizing the problems of climatic change and ozone depletion. To the general public, these problems did not exist. As a result, atmospheric scientists became the gatekeepers to the atmospheric environment, and, in turn, commanded a particular authority in American environmental politics.

Scientists’ interest in the SST’s impacts on the atmospheric environment reflected a precautionary ethos they shared in common with American environmentalists, but the debate also exposed tensions between the two communities. These tensions revolved around the two groups’ attitudes toward new technologies. Environmentalists embraced scientific developments that might help them to forge a more responsible relationship with the world around them, but they objected to the centralization, secrecy, and wastefulness of many government-sponsored “big science” projects and the technologies they engendered. To groups like the Sierra Club, the Wilderness Society, and Friends of the Earth, the SST symbolized the worst elements of the aerospace boom of the 1960s.

Stewart Brand’s Whole Earth Catalog, an oversize review of tools and technologies designed to help readers break environmentally irresponsible conventions and manage their own lives in environmentally sustainable ways, embodied environmentalists’ ambivalence about technology.204 First published in 1968, the Catalog

204 Andrew Kirk argues in Counterculture Green that Brand’s optimistic, technologically savvy form of environmentalism represented a unique and distinct wing of the environmental movement. To be sure, The

93 embraced the potential benefits of transparently produced, widely disseminated knowledge and technology. For Brand and many other environmentalists, the main problem was not the idea of technological development itself, but its administration by an exclusive central authority that failed to sufficiently account for environmental values.

Environmentalists were fascinated by new technologies, but they often distrusted the institutions that created them. The Catalog enabled individuals to capitalize on technological gains while operating outside of these institutional boundaries.205

Embedded in the Catalog’s populist, do-it-yourself message was also a critique of the inaccessible, bureaucratically controlled technologies that supported an environmentally irresponsible status quo. Environmentalists were particularly skeptical of large-scale corporate and government science and technology projects like the SST that they felt paid little heed to citizen’s input and disproportionately valued technological or material gains over the goals conservation and environmental protection.

Atmospheric scientists were far less distrustful of government, and they relied on the high-tech tools of big government science to do their jobs. NOAA and NASA were government institutions themselves, and atmospheric scientists at NCAR and elsewhere frequently collaborated in government-sponsored research. 206 The scale of the atmospheric environment required expensive, large-scale research projects. Atmospheric

Whole Earth Catalog and its acolytes were both unique and optimistic, but the Catalog, rather than defying the mainstream of environmentalism, actually helped define practical environmentalism for individuals who hoped to affect change at the personal level. Kirk argues that Brand only bought into parts of the environmental movement, but the same can be said for many environmentalists in the 1970s. For more on Brand, also see Fred Turner, From Counterculture to Cyberculture: Stewart Brand, the Whole Earth Network, and the Rise of Digital Utopianism (Chicago: University of Chicago Press, 2008). 205 Increasingly in the 1960s and early 1970s, their objections to technological development revolved around energy use. The Catalog introduced a mass audience to renewable alternative energy sources like geothermal, solar, and wind power. Those technologies that either consumed excessive amounts of energy (like the SST) or produced energy in inefficient or environmentally destructive ways came under attack. 206 For a more detailed discussion of instutions of American atmospheric science, see chapter 1.

94 scientists were eager to apply existing and developing aerospace technologies to their research into the global environment, and they expressed particular excitement about the computer and the satellite.207 They generally sympathized with the conservation goals of the environmental movement, and individual scientists contributed time, money, and expertise to environmental organizations like the Sierra Club and the Wilderness Society.

NCAR director John Firor, for example, sat on the board of Environmental Defense.208

But members of the atmospheric science community often regarded environmental organizations’ opposition to the SST as a knee-jerk reaction against technology more generally. Many agreed with Fred Singer when he chastised environmentalists for fostering “a general reaction against all technological progress and against basic science itself.”209

Not all scientists were sanguine about all technology. Many American scientists struggled to reconcile their commitments to technological development with their concerns over new technology’s potentially destructive impacts. Divisions between scientists often ran along disciplinary and institutional lines. Outspoken biological scientists—especially conservation biologists and ecologists working at universities—took the lead in warning citizens about the impacts of technology on the quality of their environment. They opposed developments made possible, and often supported, by chemists, geologists, and nuclear physicists in both government and industry. As early as 1962, Rachel Carson used the science of ecology to illuminate the

207 In its financial impacts section, SMIC considered the cost of a five year program of Supersonic flying time—roughly $30 million—noting that it was “a very small sum compared to that already spent on the development of the aircraft that fly, or will fly, in the stratosphere.” SMIC, 7. 208 Alison Peterson, “John Firor, 80, Early Voice on Environment, Is Dead,” The New York Times, November 12, 2007, http://www.nytimes.com/2007/11/12/us/12firor.html. 209 Quoted in “Editorials: SST—Some Answers, Some Questions,” Boulder Daily Camera, March 15, 1971.

95 environmental risks of cutting-edge chemical pesticides.210 In the late 1960s Washington

University biologist Barry Commoner and Stanford University entomologist and population biologists Paul Ehrlich each built on this strategy in popular books meant to expose the imminent environmental dangers associated with technological development and human population growth.211 Some scientists, like William Shurcliff and John

Edsall, objected to new technologies in their capacity as citizens rather than as scientists, although activist scientists like Commoner argued that the two roles should not—in fact could not—be separated.212

The Vietnam War also led many scientists to question the impacts of technological development. Liberal scientists, again primarily at universities, objected to conducting research used in America’s military efforts in Vietnam. At MIT, graduate students and some faculty joined in with protestors who called for a reduction in the more than $43 million per year the institution received from the Defense Department.213 On

March 4, 1969, a group of graduate students and distinguished professors held a work stoppage in order to highlight the threat posed to humankind by “the misuse of scientific

210 Rachel Carson, Silent Spring (New York: Houghton Mifflin Company, 1964). 211 Chapters 2 and 4 address Commoner and Ehrlich in greater detail, though not comprehensively. John Egan’s Barry Commoner and the Science of Survival (Cambridge: MIT Press, 2007) quite convincingly uses Commoner as a lens through which to view the American environmental movement more broadly. He demonstrates how Commoner’s dual commitments to science and environmental activism first gained him popularity, and later notoriety, as he sought to implement what Egan calls an “apparatus” for environmental stewardship consisting of popular dissent, dissemination of technical information, and a public discussion of environmental risks. A definitive biography of Ehrlich has yet to appear. 212 Shurcliff and Edsall highlight the problems of categorizing scientists in terms of their environmental advocacy and their relationship to technology in the late 1960s and early 1970s. Both men operated chiefly in their capacity as citizens rather than as professional scientists, and neither had any particular expertise in sonic booms. Nevertheless, they initially framed their opposition to the SST in terms of the biases of the government’s scientific reports, and their Harvard credentials helped give their claims a certain scientific credibility. 213 Kevles, The Physicists, 402-403.

96 and technical knowledge.”214 At Stanford, Denis Hayes—later a key coordinator of the first Earth Day—helped to lead the takeover of an applied electronics laboratory conducting classified defense research.215 According to a Physics Today study conducted in 1972, four out of five physicists surveyed disapproved of conducting classified research on university campuses, and opposed Nixon’s action in Vietnam.216

Objections to the Vietnam War often entwined with scientists’ environmental concerns. The AAAS, America’s largest scientific society, passed a series of resolutions between 1965 and 1972 that encouraged a peaceful resolution to the Vietnam War, called for a study of the effects of America’s use of environmental modification techniques in the war effort, and finally denounced both the use of biological and chemical weapons and the war itself.217 Scientists’ concern for the health of the environment provided a platform from which to criticize other scientists’ roles in creating technology that served the war effort.

The generation of atmospheric scientists that had overseen the tremendous growth of their discipline during the aerospace boom of the1950s and ‘60s—Revelle, Roberts,

214 An edited overview of the speeches and panels conducted during the work stoppage can be found in Jonathan Allen (ed)., March 4: Scientists, Students, and Society (Cambridge: MIT Press, 1970), xxxii. 215 Interview with Denis Hayes, June 2, 2009, in Seattle, WA with the author. 216 “Survey Finds Physicists on the Left,” Physics Today, XXV (October, 1972), 61-61, cited in Kevles, The Physicists, 405. Kevles accepts the study’s conclusion that American physicists “constituted the most politically liberal group” among American natural scientists. Certainly, many elite physicists—the leaders in their fields—held liberal views, and were outspoken in their opposition to nuclear proliferation and the war in Vietnam. But the study, and Kevles’s book, treats particle physicists as really the only physicists that matter in American science. His work largely ignores atmospheric physicists, who on the whole were perhaps no less “liberal” in the sense that they supported social freedom, international cooperation, and a greater degree of government regulation, but certainly had more qualms with the anti-establishment bent of some physical and biological scientists. 217 For “AAAS Resolution: Study of Ecological Effects of Herbicides,” “AAAS Resolution: Use of Herbicides in Vietnam: Call for Field Study,” “AAAS Resolution: Commending U.S. Government Phase-Out of Herbicides in Vietnam,” “AAAS Resolution: Assessment of the Ecological Effects of U.S. Activities in Vietnam,” and “AAAS Resolution: Appeal for Cessation of Hostilities in Vietnam,” see “About AAAS: History & Archives,” http://archives.aaas.org/docs/resolutions.php.

97 Kellogg, and others—genuinely worried about the impacts of technological change on both humans and their environments, but they understood environmental issues in utilitarian terms. On the whole, the discipline’s leaders were less radical than some of their counterparts in particle physics, biology, and environmental science in 1970.218

They saw environmental problems as just that—problems—and as scientists, they sought to identify and solve these problems.219 Most atmospheric scientists investigated the

SST’s effects on the atmosphere not as a way to undercut the project, but as the first step in a process of making the plane better. The second step, they believed, would be to help mitigate these effects in a better design for the SST of the future.220

In December of 1970, in the midst of the debate over the SST, 34 scientists from institutions around the country signed a statement by Presidential Science Advisor

Edward E. David, Jr. that supported further funding for the SST and articulated many atmospheric scientists’ views on technological development. “Our society must not suppress technological advances,” David wrote,

“but through research, development, and experimentation make sure that those

advances are obtained without undesired side effects. Instead of canceling work

218 As I will discuss in the next chapter, these types of blanket statements about the politics of atmospheric scientists can be very misleading. Even characterizing sub-disciplines of atmospheric science can be tricky, and during the 1970s, the politics of science, as well as scientists’ politics, were very much in flux. Nevertheless, it is useful to compare leaders in atmospheric science with other vocal scientists in order to understand their relationship to other environmentalists and other scientists. 219 In this respect, atmospheric scientists had more in common with the Progressive Era conservationists and environmental managers that Sam Hays describes in his work on the early 20th Century conservation movement, Conservation and the Gospel of Efficiency, than they did with the largely defensive environmentalists Hays presents in Beauty, Health, and Permanence. This comparison will be investigated further in discussions of climate scientists and the world food crisis in the following chapter. Samuel P. Hays, Conservation and the Gospel of Efficiency: The Progressive Conservation Movement, 1890-1920 (Cambridge: Harvard University Press). 220 NASA, for example, began a research program at the Lewis Research Center focusing on the possibility of lower-emission jet engines. A later study, the Climate Impacts Assessment Report, emphasized the potential of “future technology” developed at NASA to reduce ozone depletion to a “minimally detectable” level, or .5%. Conway, Atmospheric Science at NASA, 208.

98 on the SST, we should mount a vigorous program of experimentation aimed not

only at solving the technical problems of economic supersonic transportation but

also assuring no undesirable effects.”221

Even when they did share environmentalists’ views on the environmental impacts of technological development, atmospheric scientists had difficulty reconciling environmentalists’ direct political activism with their own ideals of political neutrality and scientific objectivity. They sought to contain their professional concerns over atmospheric pollution and anthropogenic climate change within the community-defined boundaries of “good science.” The “ist” in “scientist” was meant to supersede the “isms” of political or ideological commitments. Ultimately, scientists’ commitments to the ideals of “good science” limited the nature and extent of scientists’ environmental advocacy.

The concepts of neutrality and objectivity are not interchangeable. As Robert

Proctor argues in Value-Free Science?, scientists employ the concept of “value neutral” or “value-free” science as a way to question the extent to which science serves a political or ideological end. “Neutrality,” he writes, “refers to whether science takes a stand.”222

221 The heterogeneity of the discipline as a whole extended to its leaders, who even as individuals often defied categorization. NCAR founder Walt Roberts is a good example. Perhaps more than anyone at NCAR, Roberts made a strong commitment to using science for the benefit of society. He served as President of the AAAS in 1968, and presided over that society’s resolution on a study of the ecological effects of herbicides in Vietnam. A New Deal Democrat and prominent scientist, the FBI at one point suspected he had communist leanings, and kept a file on him through the late 1940s and early 1950s. But like Singer and Kellogg, he also harbored a strong commitment to science and its institutions, and he signed David’s statement in 1970. As I discuss in the following chapters, however, his commitment to the social relevancy of science remained, and even after his ouster at NCAR, he would help that institution become a leader in studying climatic change. Walter Orr Roberts FBI File (FOIA), National Center for Atmospheric Research Archives; Department of Transportation Press Release, December 5, 1970, SST Testimony File, William W. Kellogg Files, NCAR Archives, Boulder, Colorado. 222 Robert Proctor, Value-Free Science?: Purity and Power in Modern Knowledge (Cambridge: Harvard University Press, 1991), 10. The concepts of neutrality and objectivity have stood at the center of the History of Science for decades, and innumerable scholars have taken up the subject in their research. I point to Proctor’s work here because I find it particularly clear and accessible. Proctor also historicizes

99 In Proctor’s formulation, neutrality need not have anything to do with concept of

“objectivity,” which he defines as “whether science merits certain claims to reliability.”223

When scientists contest opponents’ political neutrality, however, they often do so to undermine their credibility, and they frame neutrality as a pre-requisite for true objectivity. They assume that science conducted for a specific purpose or influenced by outside considerations cannot ultimately be as reliable as science conducted for its own sake.224 As a result, the terms—and the concepts—are often conflated.

The concepts of “neutrality” and “objectivity” are hardly absolute. Most scientists harbor commitments to both ideals, but scientists themselves define the boundaries of these concepts, and often these definitions are only implicit until the boundaries are transgressed.225 Atmospheric scientists studying the SST, for example, rarely used the word “neutrality” or “objectivity” when discussing atmospheric research, but they employed the concepts frequently in their commendations of “good science” and

these concepts in a way that narrower studies could benefit from. Environmental historians, too, have taken up the subject. In Nature’s Experts, Bocking relates these concepts to the dual character of scientific authority in environmental affairs. On one hand, he argues, scientists create authoritative knowledge on environmental issues like water pollution or landscape management. On the other hand, however, adversaries in environmental disputes frequently attempt to undermine opponents’ positions by challenging the neutrality—and thereby the authority—of their science. Among other things, Sam Hays discusses the relationships between competing groups of experts and environmental advocacy groups in Beauty, Health, and Permanence. For more on objectivity and neutrality, see Helen E. Longino, Science as Social Knowledge: Values and Objectivity in Scientific Inquiry (Princeton: Princeton University Press, 1989), Karl Popper, ObjectiveKnowledge (Oxford: Oxford University Press, 1971). For objectivity in the profession of history, see Peter Novick, That Noble Dream: The “Objectivity Question” and the American Historical Profession (Cambridge, England: Cambridge University Press, 1988). 223 Proctor, Value-Free Science?, 10. 224 Proctor makes the distinction between objectivity and neutrality clearer when he describes objective, value-laden science. Science can have a particular political or economic motivation (i.e., a set of values), but still be conducted in compliance with the standards of a discipline—standards that ensure that the results are reliable. Oil company geologists, he points out, know more about oil-bearing shale than other rocks, but their goals as employees of oil companies make this knowledge no less reliable. Proctor, Value- Free Science?, 10. 225 Scientists hold neutrality and objectivity as professional ideals in their day-to-day work, but these concepts are rarely employed outside of the context of a scientific controversy. During controversies, opposing groups begin to define neutrality as they use competing neutrality claims as a tool to undermine their opponents’ expertise.

100 the “good scientist,” as well as in their condemnations of politically-motivated science and “emotionalism.”

Scientists whose work most closely conforms to an existing consensus on an issue often have stronger claims to political neutrality, especially when this consensus implicitly supports a dominant political position.226 Some environmentalists cautiously pointed out that atmospheric scientists developed a consensus on the SST largely within government institutions that had political and financial interests in a continued program of SST research.227 Leaders like Singer and Kellogg at governmental and quasi- governmental institutions dominated the field, and their positions on the SST hewed to a consensus established by the SCEP and SMIC studies, as well as by other scientists at

NCAR, NASA, and elsewhere. As the dominant specialists in the field, Kellogg and

Singer thus helped set the boundaries of “good science” in atmospheric research.

When outspoken opponents of the SST like James MacDonald and Harold

Johnston used their science to make an explicit case against the supersonic, they transgressed these boundaries, giving supposedly “neutral” scientists grounds to question the validity of their research. Kellogg’s review of MacDonald’s written testimony on the

SST in the spring 1971, for example, complained that the document was “angry and finger-wagging, clouding the real issues.”228 “He is a good scientist,” Kellogg acknowledged, “and most (though not all) of his facts are correct—but he presents the

226 Bocking, Nature’s Experts. 227 As Laurence Moss of the Sierra Club and Gary Soucie of Friends of the Earth pointed out, the individuals called to testify before Congress on the SST as representatives of a consensus position had political motives of their own. “Without slurring the scientists reputations,” the Washington Post reported, “conservationists noted that two [of the expert witnesses, Kellogg and Singer] were government employees, and the third [Leo L. Beranek] frequently a government contractor.” David Hoffman, “Three Scientists, Hedging Slightly, Give the SST a Clean Bill of Health,” Washington Post, March 4, 1971. 228 William Kellogg comments on MacDonald draft (no date), SST Testimony File, William W. Kellogg Files.

101 case with too much emotion and a rather transparent hostility to the whole idea of the

SST.”229 Kellogg didn’t immediately object to the details of MacDonald’s work—in fact, he called for a careful reconsideration of MacDonald’s main points.230 Instead, he undermined the study’s validity by highlighting its apparent bias. Scientists like

MacDonald who hoped to highlight the specific threats that the SST posed to the atmospheric environment expressed their political opposition only at the risk of damaging their professional reputations as scientists.231

Kellogg and his colleagues genuinely objected to MacDonald and Johnston’s

“politicization” of atmospheric science, but they also used neutrality claims as a way to protect their own status in battles between institutions and sub-disciplines. The controversy over the SST reflected real divisions between scientists from different disciplines working with different methodologies at different institutions. Atmospheric scientists criticized Johnston in particular for attempting to apply his expertise in ground- level chemistry to the unique conditions of the stratosphere, where they believed effects

of NOx on Ozone would be dampened by the extremely low temperature and density of the air. Johnston’s numbers worked within the confines of his Berkeley laboratory, but scientists at NCAR and NOAA argued that his methods failed to account for the complexities of the atmosphere. They believed that only their own sub- disciplines—primarily atmospheric chemistry and atmospheric physics—could appropriately address the particularities of the SST problem.

229 Ibid. 230 Ibid. 231 Kellogg didn’t discount MacDonald by any means because of his bias; he simply found it unconvincing on the particular issue of the SST. “Jim MacDonald,” he noted, “would not have raised these worries without having some good arguments to back him up.” Ibid.

102 Scientific disagreements blended with disagreements about how and where the science ought to be conducted, and Kellogg and his colleagues used challenges to

Johnston’s political neutrality as a trump card. Johnston’s work undermined the political status quo by emphasizing the SST’s damaging impact on the atmosphere, but it also threatened NCAR and NOAA scientists’ primacy as atmospheric experts. By combining their disciplinary criticisms of Johnston’s methodology with attacks on his political neutrality and, by implication, his objectivity, atmospheric scientists at NCAR and

NOAA effectively undermined Johnston’s authority as an expert on the atmosphere while reaffirming their own primacy in the field.232

Direct political advocacy like Johnston’s typically transcended the boundaries of

“good science,” but atmospheric scientists frequently used environmental concerns to advocate on behalf of their institutions and disciplines within the larger scientific community. Like the concept of neutrality itself, the line between political and scientific

232 During CIAP, Johnston was able to convince many atmospheric scientists that his evidence did in fact show that the SST was a significant environmental threat, but by this time, the political issue had largely passed. Johnston also worked to integrate the particularities of the atmosphere into his science by collaborating with atmospheric scientists, thereby eroding some of the disciplinary divisions that fostered the earlier rift. It is tempting to be cynical about relationships between scientists in the midst of scientific controversy, and literature on scientific controversy only feeds that cynicism. In Science in Action (Cambridge: Harvard University Press, 1987), Bruno Latour presents science, or at least “technoscience,” as an essentially agonistic endeavor, wherein scientists’ primary concerns are for power or influence rather than any search for an elusive truth about the natural world. To an extent, this viewpoint is useful in attempting to discover the motivations of scientists—particularly environmental scientists—as they variously support and thwart environmentalists’ efforts to use science in their social and political campaigns. Hays, though he doesn’t cite Latour, essentially presents a similarly cynical view of environmental scientists in his discussion of the internal politics of science, wherein scientists from one discipline strive to establish their authority as experts in certain fields by using claims of bias to undermine the expertise of scientists from different disciplines. The SST debate supports this framework, but the idea that science is agonistic should not be taken too far. It is certainly not purely agonistic; in fact, it is largely collaborative, and many scientists relish pointing to examples where their own science has been superseded by other scientists’ further research. On a day-to-day basis, moreover, scientists’ primary concerns are by and large not political at all, but technical. Though scientists engage in disciplinary and institutional politics, especially during controversies, it is all too easy to ignore what sometimes seems like a naïve but is in fact a real commitment to “good science.” In that sense they behave like many other professionals who, though they have professional political concerns, spend the bulk of their time and energy trying to excel at the tasks they are paid to perform. Letter from Harold Johnston to William Kellogg, October 2, 1972, William W. Kellogg Files.

103 advocacy was fuzzy, and institutional leaders often had to balance their commitment to neutrality against the obligation to promote the interests of their science. Funding was a key concern. Through executive-level organizations, the federal government funded the majority of atmospheric science conducted in the U.S. NCAR, technically a non- governmental research center, relied especially heavily on the federally controlled

National Science Foundation for funding; NOAA, a government institution, leaned on its parent agency, the Department of Commerce.233 NASA was its own federal agency.

Each of these agencies, in turn, ultimately depended on Congress and the executive

Office of Management and Budget to approve their budgets. Congress increasingly demanded that the nation’s scientific institutions directly serve the “national interest,” and in the early 1970s, both agencies began to put pressure on atmospheric scientists to make their basic research more relevant to contemporary social and political issues.234

They hoped to see tax dollars spent on practical solutions to the complex problems of the

1970s. The political salience of environmental issues, along with many scientists’ genuine concern for humans’ impact on the global atmosphere, prompted some scientists to frame their research on the atmosphere in terms of environmental degradation. In the case of the SST, the trick was to convince Congress and the administrators at the NSF and in the Department of Commerce that the plane presented enough of a potential threat

233 According to NSF administrators, half of the foundation’s overall expenditures on the atmospheric sciences went to NCAR. As a subsidiary of the Department of Commerce, NOAA’s funding was more secure, but so too was its mission to conduct socially applicable research extremely clear. National Science Foundation Advisory Panel for Atmospheric Sciences, “Report on the National Center for Atmospheric Research,” November , 1970, NCAR Archives. 234 Ibid.

104 to merit expensive atmospheric research, but without “overstating” the case in a way that might unduly threaten the SST—and its budget.235

Funding was essential for atmospheric scientists to continue their day-to-day research, but Kellogg and his colleagues sought more than just funding. They also hoped to gain influence in the national and international organizations that helped determine international research priorities and set governments’ scientific agendas. These organizations included governmental agencies like the NSF and Department of

Commerce, independent non-governmental organizations like the American Association for the Advancement of Science (AAAS), international non-governmental organizations like the International Council of Scientific Unions (ICSU), and United Nations agencies like the World Meteorological Organization (WMO). Later, this list also included the

U.N. Environment Programme (UNEP). Like the NSF, these organizations responded to what they saw as society’s pressing scientific needs, and again atmospheric scientists and their colleagues studying global systems presented their research in environmental terms.

The SCEP and SMIC reports provide a neat example of scientists’ efforts to promote the interests of their disciplines. The reports’ authors framed global environmental research in terms of a global environmental crisis that required action, but their recommendations revolved almost entirely around the functioning of science itself.

The primary audience of the SMIC report, it’s authors noted, was the scientific community. “Our secondary audience,” it continued,

“includes those national officials whose decisions govern the allocation of

resources to support scientific programs and the directors of those institutes and

235 Johnston to William Kellogg, October 2, 1972, William W. Kellogg Files.

105 laboratories where programs must be initiated, expanded, or modified to

implement these recommendations.”236

One of the chief goals of the study was to convince this secondary audience to grant atmospheric and climatic science “a sufficiently high priority to justify allocation of the resources needed”—that is, to present threats to the atmospheric environment as dire enough to merit studying.237 Both SCEP and SMIC played up the dangerous uncertainties of interfering with global environmental systems, but their measured conclusions led inevitably to recommendations for more research rather than real domestic public policy changes, at least in the short term.

The distinction between promoting scientific research and advocating public policy was not always clear, especially at the international level. Atmospheric scientists shied away from domestic environmental policy proscriptions, but they actively promoted a movement within the United Nations to reevaluate the relative priorities of development and environmental protection. In 1968, the United Nations began planning for a 1972 Conference on the Human Environment, and both SCEP and SMIC were intended as supporting documents for the Conference. The studies outlined scientists’ concerns over a new type of international environmental issue for the United Nations to grapple with. The SCEP study “focused on environmental problems whose cumulative effects on ecological systems are so large and prevalent that they have worldwide significance.”238 These were not bread-and-butter environmental issues like water quality and air pollution involving “the direct health effects of pollution on man,” the study

236 SMIC, 4. 237 SMIC, 4. 238 SCEP, 5.

106 contended.239 Rather, these were long-term, large-scale environmental changes only recently identified by scientists. The environmental impacts of these changes as yet fit into no specific national or international regulatory frameworks.240 “Should preventative or remedial action be necessary,” SMIC noted the following year, “it will almost certainly require cooperation among the nations of the world.”241 The studies were developed to provide a scientific consensus for this international environmental decision- making process, but the studies’ recommendation also assured scientists themselves a place in international political discussions about development and natural resources. “It is hoped,” SMIC continued, “that this consensus will provide an important input into planning or the 1972 United Nations Conference on the Human Environment and for numerous other national and international activities.”242

In contrast to environmentalists, scientists rarely pitched their concerns over the atmosphere directly to the public. American environmentalists appealed to middle-class

Americans for support in protecting the nation’s wilderness areas and natural resources; activist scientists appealed to an international governmental elite for a change in high- level science policy in order to monitor and better understand the global problem. If atmospheric science achieved a higher priority in an international scientific agenda, they believed, the national and international bureaucrats responsible for setting environmental policy would receive better information and therefore make better decisions. “Our third audience,” the SMIC authors wrote, “is comprised of those international organizations whose encouragement was vital in carrying out this Study…and whose response to our

239 SCEP, 5. 240 SCEP, 5. 241 SMIC, xv. 242 SMIC, xv.

107 recommendations…will be so important in achieving early action at all levels.”

Specifically, the report named the WMO, the ICSU, the Scientific Committee on

Problems of the Environment (SCOPE), and the Joint Organizing Committee of the

Global Atmospheric Research Program (JOC/GARP) as key conduits between atmospheric scientists and the upcoming U.N. Conference.243 The image of the Earth on

SMIC’s cover and the Sanskrit prayer in the frontispiece conjured a popular ethos of environmental protection, but nowhere in either SMIC or SCEP did the authors attempt to mobilize the public at large.

Conclusion

There is little evidence that concerns over the atmosphere played more than a marginal role in Congress’s decision to deny funding for the American Supersonic in

1971, but the debate over the ill-fated program introduced large-scale atmospheric change as a mainstream environmental issue for the first time. It also brought atmospheric scientists a greater role in American environmental politics, though not without some ambivalence. The global scale of the atmosphere and the complexity of its impacts on natural and human systems distinguished atmospheric and climatic change from other mainstream environmental issues. Only atmospheric scientists had the high-tech tools and disciplinary expertise to recognize threats to the Earth’s atmosphere, and as informational gatekeepers during the SST controversy and afterwards, they were forced to act as the primary advocates for this newly threatened environment. If atmospheric scientists shared with their counterparts at environmental organizations a common

243 For more on GARP, the WMO, and the ICSU, see chapter 1.

108 precautionary ethos, however, scientists made atypical environmentalists at best. Their particular forms of advocacy reflected their professional values. As scientific professionals, atmospheric scientists were meant to value the community-defined ideals of “good science” over their personal political commitments, and they shared a faith in—and dependence on—the technological advances of the aerospace era that contrasted with many environmentalists’ distrust of large-scale high-tech development.

Nevertheless, atmospheric scientists took the lead in promoting atmospheric change as a major environmental issue in the 1970s. In particular, they focused their efforts on influencing the agenda of international scientific and environmental organizations leading up to the United Nations Conference on the Human Environment in 1972.

In part, differences in atmospheric scientists’ and professional environmentalists’ forms of advocacy reflected differences between the two groups’ institutions.

Environmental organization like the Sierra Club and Wilderness Society operated under a central political mission, supported by members and donors, to advocate for conservation and the protection nature, wilderness, and the broader environment. They were largely democratic, and often engaged in collective political campaigns on common issues. They were—and still are—advocacy groups. Scientists, by contrast, worked in competitive—and sometimes competing—institutions with the primary and overriding objective of producing more and better knowledge. Where they existed, hierarchies within institutions like NCAR, NASA and NOAA were chiefly based on merit and seniority, and these institutions answered to bureaucratic agencies. To the extent that the larger atmospheric community had a central collective political motivation, it was the promotion of science itself. Atmospheric scientists framed their research in terms of its

109 potential social applications—applications that included environmental assessments and resource management. When it came down to policy, however, the scientific community’s first priority was to support programs that allocated more money for scientific research.

Markedly absent from scientists’ appeals to international scientific elites for more money and better cooperation in atmospheric research was a concerted effort to involve the American public in their efforts to study and protect the global atmosphere.

America’s mainstream environmental organizations relied on a broad base of grass-roots support from the middle class in their public campaigns to influence domestic environmental policy. Scientists instead sought to gain influence among high-level bureaucrats and government officials at organizations associated with the United Nations who they hoped would sponsor extensive atmospheric research that would ultimately underpin sound environmental policies. The SCEP and SMIC reports did gain a popular audience, and a related work, The Limits to Growth, quickly became an international bestseller thanks in part to its release during the 1972 U.N. Conference on the Human

Environment. But while these works certainly had polemical elements, they did not call on the public at large to take action, nor did they point to any specific sectors in business or government for everyday people to rally around or against.

Atmospheric scientists were themselves largely a middle-class group, and many among them participated in the activities of America’s environmental organizations. But as a group they actively sought to divorce their personal values from their professional opinions, and this shaped their approach to the atmospheric environment. They couched their discussions of potential threats to the global atmosphere in the equivocal language

110 of scientific uncertainty, and their calls for action largely targeted those government officials and scientific elites who controlled the budgets and agendas of science itself.

Scientists’ commitments to objectivity and political neutrality continued to limit the extent of their advocacy in the early 1970s. Their cautious scientific concerns about atmospheric change failed to capture the interest of America’s major environmental organizations in a significant way until the 1980s, and even then, scientists directed their appeals more toward governments and other scientists than toward the public at large. As a result, scientific concern over changes to the Earth’s atmosphere and climate initiated by the SST grew into a distinct—but not altogether separate—form of environmental activism guided more by the professional values of science than by the middle-class consumer values at the heart of mainstream American environmentalism. Only as environmentalism itself became more global and more scientific did atmospheric change become a central concern of America’s professional environmentalists.

111 Chapter 3 Systems Science, the Stockholm Conference, and the Making of the Global Environment

During the controversy over the American SST, atmospheric scientists began to exercise unprecedented influence in American environmental politics. But for many of these scientists, especially those involved in the 1970 Study of Critical Environmental

Problems and the 1971 Study of Man’s Impact on Climate, the political value of atmospheric research went well beyond the domestic debate over supersonics.244 These scientists understood climatic and atmospheric change in terms of a larger global environmental crisis—a crisis that, as experts in one of the Earth’s most complex global environmental systems, they hoped to study and help to manage. With working groups on climate, energy, agricultural waste, industrial pollution, and global ecological change, the SCEP and SMIC studies laid out a comprehensive scientific framework for understanding the Earth’s threatened large-scale environmental systems. Focused on environmental problems that were global in scale—problems like atmospheric and climatic change—SCEP and SMIC sought to influence a developing conversation about international environmental governance that would come to a head at the United Nations

Conference on the Human Environment in Stockholm, Sweden in June of 1972. As the

Stockholm Conference approached, however, scientists quickly discovered that theirs was only one of many contested definitions of the global environment, and that ultimately competitive national politics, rather than cooperative global science, would determine how the Earth’s complex global systems would be governed.

244 Man’s Impact on the Global Environment: Assessment and Recommendations for Action, Report on the Study of Critical Environmental Problems (Cambridge: MIT Press, 1970); Inadvertent Climate Modification: Report of the Study of Man’s Impact on Climate (SMIC) (Cambridge: MIT Press, 1971), xv, xvii. Hereafter “SCEP” and “SMIC,” respectively.

112 The U.N. Conference on the Human Environment was the seminal event in the making of the global environment as we know it today. An international conservation movement focused on both natural resources and wilderness areas dated back at least to the end of the First World War, but only in the early 1970s did the idea of a single, unified global entity in need of management and protection gain prominence in international environmental politics.245 Both scientists and environmentalists recognized that certain environmental processes, particularly those involving the oceans and the atmosphere, actually occurred at a global scale. These global processes, unfolding in a global space, defied rather than defined the typically local scale of most individuals’ relationships with natural world.246 Recognizing the limitations of contemporary domestic environmental politics, atmospheric scientists—and systems scientists more generally—sought to incorporate their large-scale, interdisciplinary approach to environmental problems into the cooperative global mission of the United Nations. For scientists and some U.N. leaders, the Stockholm Conference thus became the center of an effort to promote and codify a broader understanding of humanity’s global, collective relationship with nature.

The concept of the “global environment” that these scientists and diplomats promoted at Stockholm reflected the related natural phenomena, scientific methodologies, and political ideals behind its creation. In SCEP and SMIC, atmospheric and other systems scientists focused mainly on building more robust and cooperative global networks for scientific research, but their focus on the interdependence of the

245 For more on the early international environmental movement, see Robert McCormick, Reclaiming Paradise: The Global Environmental Movement (Bloomington: Indiana University Press, 1989). 246 See Dipesh Chakrabarty in “The Climate of History: Four Theses,” Critical Inquiry 35 (Winter, 2009): 203, 206-7.

113 Earth’s large-scale systems also implicitly—and sometimes explicitly—supported continuing efforts within the United Nations to foster global political cooperation on environmental and non-environmental issues alike. Overconsuming, polluting, or otherwise acting irresponsibly on the part of any single nation or group of people, scientists argued, affected the health of the entire global system. Moreover, they contended, any attempt to deal piecemeal with these global problems without regard to the complex feedback mechanisms built into the Earth’s large-scale social, environmental, and technical systems would at best fail—at worst, it might catastrophically backfire. The bilateral competition of the Cold War that had come to define international relations since the late 1940s had no place in this complex, systems- oriented world. Only through cooperative international institutions like the United

Nations and non-governmental organizations like the International Council of Scientific

Unions could the international community hope to grapple with a so-called monde problématique that was at once social, political, economic, and environmental.

At Stockholm, however, this vision of scientific and political unity quickly ran into the realities of an international political system dominated by national and regional interests and characterized not by cooperation, but by rivalry and competition.

International preparations for the Conference made it clear that the politics of the world’s threatened global spaces would reflect the concerns of constituencies tied to local, regional, and national geographies rather that the utopian ideals of scientists and well- meaning U.N. bureaucrats. Members of the international political community sought to reshape the image of the global environment in ways that reflected their particular

114 political concerns, and these concerns were often neither necessarily “global” nor

“environmental” in the way First World scientists and international leaders had imagined.

Two main geopolitical issues dominated the nascent politics of the global environment. First and foremost, U.N. member states understood environmental concerns in the context of international development. Less developed countries (LDCs) not only disagreed with representatives from the developed world over the relative priorities of environmental protection and economic development, they also identified a separate set of environmental problems associated with poverty that they pushed as central to any international regime of environmental regulation. Second, despite U.N. leaders’ calls for a “new globalism” inspired by common environmental challenges, the entire conversation played out against the backdrop of the Cold War. The Soviet Union, in an ultimately unsuccessful attempt to gain parity between East and West within the U.N., boycotted the Conference. Powerful non-nuclear states like Sweden,

Japan, and China used the international conversation about environmental degradation as a forum for attacking the United States for its controversial positions on disarmament and the war in Vietnam. Conference delegates validated their criticisms by expanding their definitions of environmental degradation, but in the end these conversations had less to do with the environment than they did with larger battles over ideology and power.

Ultimately, the U.N. Conference on the Human Environment established an impressive framework for international environmental governance, but it was a framework characterized less by a new ethos of cooperation and interdependence than by the competing ideologies and international rivalries that dominated international politics more broadly. It was a framework, moreover, built around issues like natural resource

115 allocation, land use, pesticides, and waste management, that though endemic across the

Earth were issues rooted in specific local and regional landscapes that housed definite political constituents. It left little room for the one truly global environmental issue that would haltingly emerge as the most important international environmental problem the world would face in the decades to come—the issue of climate change.

Climate, Systems Science, and the Global Environment

From the beginning, U.N. leaders recognized that any kind of regime of international environmental governance would necessarily include a broad program of scientific research and environmental monitoring. If the Earth was, as U.S. Ambassador to the United Nations Adlai Stevenson pronounced in 1965, “a little spaceship, dependent on its vulnerable reserves of air and soil,” it was up to scientists to find out exactly what kind of shape the ship was in before politicians could make a plan to save it.247

In the run-up to the 1972 U.N. Conference on the Human Environment, scientists produced four major collaborative reports on the Earth’s large-scale environmental systems. They were the reports of the Study of Critical Environmental Problems (SCEP)

247 There is some disagreement over the coining of the phrase “Spaceship Earth,” but it began to appear regularly in the mid to late 1960s (after Stevenson’s speech), first in popular economic literature through Barbara Ward’s book of that title, and through Kenneth Boulding’s 1966 essay “The Economics of the Coming Spaceship Earth.” With images of the Earth from space produced by NASA satellites to drive home the point, it became part of common usage by the end of the decade. The metaphor of Earth as an isolated ship moving through space is not a new one, however. The image appears as early as the 19th century in Henry George’s Progress and Poverty: an inquiry into the cause of industrial depressions, and of increase of want with increase of wealth. The remedy (New York: D. Appleton, 1880), although this image was not “environmental” in the way that Stevenson, Ward, and Boulding presented it. See Adlai E. Stevenson, “Strengthening the International Development Institutions,” speech before the United Nations Economic and Social Council, Geneva, Switzerland, July 9, 1965, as reprinted online at www.adlaitoday.org/article.php?id=6; Barbara Ward, Spaceship Earth (New York: Columbia University Press, 1968); Kenneth Boulding, “The Economics of the Coming Spaceship Earth,” Environmental Quality in a Growing Economy, Essays from the Sixth RFF Forum, edited by Henry Jarrett (Baltimore, MD.: Johns Hopkins Press, 1966): 3-14. An early version of Boulding’s paper, entitled “The Earth as a Spaceship,” and dated May 10, 1965, also exists in the Kenneth E. Boulding Papers, Archives (Box # 38), University of Colorado at Boulder Libraries.

116 and its follow-up conference, the Study of Man’s Impact on Climate (SMIC), a report by the International Council of Scientific Union’s “Scientific Committee on Problems of the

Environment,” entitled Global Environmental Monitoring, and The Limits to Growth, a book produced by a group called the Club of Rome. These studies were developed more or less concurrently between 1968 and 1972, and they overlapped significantly in subject matter and in personnel. They all advocated a holistic, interdisciplinary “systems science” approach to the global environment that their authors believed would, if implemented, lead to rational and effective environmental policy-making at the international level. With the exception of Limits to Growth, each of these studies was undertaken explicitly for the 1972 U.N. Conference on the Human Environment, and, again with the exception of Limits to Growth, each included a significant number of atmospheric scientists and was characterized by an overriding concern for atmospheric and climatic change.

Systems science provided the methodological foundation for the holistic approach advocated by the SCEP, SMIC, and SCOPE groups. A general term, “systems science” describes a broad and multifaceted movement within the scientific community after

WWII to use digital computers to numerically simulate large-scale, complex, nonlinear phenomena.248 Rooted in one of the foundational fields of systems science, numerical weather modeling, atmospheric science was by itself a form of systems science. But as

Paul Edwards argues in his essay, “The World in a Machine,” SCEP and SMIC represented more than just a continuation of the legacy of weather modeling. The intersection of atmospheric modeling with the management-focused science of systems

248 For a comprehensive (and highly critical) review of the foundations of systems science, see Robert Lilienfeld, The Rise of Systems Theory: An Ideological Analysis (New York: John Wiley & Sons, 1978).

117 dynamics in the early 1970s, he argues, marked a turning point in the way scientists in general approached the global environment.249

Systems dynamics, or the study of the complex organizations over time, was in large part the brainchild of a computer engineer turned management science expert named Jay Forrester.250 Forrester began his career in systems management during the

Second World War at the Servomechanics Laboratory at MIT, where he worked with the

Office of Naval Research to help interpret and partially automate Cold War strategic defense systems. Forrester stressed the nonlinear informational feedback loops that he believed drove the behavior of all complex organized entities.251 Complex systems, he

249 Paul Edwards argues that these two avenues of scientific inquiry—numerical models of weather and climate and world dynamics models of the earth’s large-scale systems—laid the groundwork for today’s concept of a knowable, manageable “world.” But while both of these disciplinary projects addressed the Earth as a global whole, neither of them would have necessarily been “environmental” had it not been for their relationship to a long-developing third application of systems science: ecosystems ecology. As Joel Hagan argues in An Entangled Bank: The Origins of Ecosystems Ecology (New Brunswick, NJ: Rutgers University Press, 1992), the concept of biological interdependence is essentially as old as biological science itself, but only after WWII did ecologists begin to try to understand the relationships between organisms as parts of broader organizations of the living and non-living parts of a particular environment, or ecosystems. In 1964, Eugene Odum coined the term “Systems Ecology,” which, according to Hagan, at once referred to disciplinarily specific efforts to model discrete ecosystems using digital computers and to a more general commitment to “the philosophical core of [Arthur] Tansley’s ecosystem concept”—namely, the “belief that nature is composed of innumerable, partially overlapping systems. As Hagan, Stephen Bocking, and others have since noted, the concepts of equilibrium, cooperation, and integration implicit in the systems approach to ecology became, in simplified form, the popular scientific foundations for 1960s American environmentalism. Hagan, An Entangled Bank, 131; Stephen Bocking, Nature’s Experts: Science, Politics, and the Environment (New Brunswick: Rutgers University Press, 2006); Eugene P. Odum, “The New Ecology,” BioScience 14, no. 7 (July, 1964), 14-16. For more on the specifics of the relationship between systems dynamics models and numerical weather models, see Edwards, “The World in a Machine,” in Systems, Experts, and Computers: The Systems Approach in Management and Engineering, World War II and After, edited by Agatha Hughes and Thomas Hughes (Cambridge: MIT Press, 2000), 221-253. For more on the term “climate science,” see Chapter 4. 250 Forrester actually only christened the methodology outlined in his Industrial Dynamics (Cambridge: MIT Press, 1961) and Urban Dynamics (Cambridge: MIT Press, 1969) as “Systems Dynamics” in 1970, when it became clear that the methodology could in fact be generalized to fit a wide variety of world-scale systems problems—the monde problèmatique, as the Club of Rome called it. For more on Forrester and the Club of Rome, see chapters three , four, and five of Fernando Elichirigoity, Planet Management: Limits to Growth, Computer Simulation, and the Emergence of Global Spaces (Evanston, IL: Northwestern University Press, 1999); see also Edwards, “The World in a Machine.” 251 Edwards, “The World in a Machine,” 237-8.

118 argued, operated through “a multiplicity of interacting feedback loops.”252 The greater the scale of a system, the less intuitive these relationships became.

The key to interpreting these non-intuitive feedback loops, he contended, was in computer models.253 Forrester sought to use computers to simulate the nonlinear relationships of complex systems numerically, and thereby provide a framework for understanding how to improve management practices, first within the U.S. Department of

Defense, and then later within the sciences of industrial management and urban development. Throughout his career, Forrester’s interests continued to grow in scale, until by the late 1960s, he and his colleagues at MIT began to address even the largest economic, political, social, and environmental systems as subsystems within an even larger holistic world system. The science that described this system represented the logical extreme of systems dynamics: “world dynamics.”

The most famous application of Forrester’s science of world dynamics—and the most controversial—was The Limits to Growth, a study published in the spring of 1972 by an elite international group of scientists, industrialists, businesspeople, diplomats, and leaders of civil society called the Club of Rome. Founded in 1968 by Italian industrialist

Aurelio Peccei, the group dedicated itself to studying the genesis of, and possible solutions to, humanity’s common and persistent problems: “poverty in the midst of plenty; degradation of the environment; loss of faith in institutions; uncontrolled urban spread; insecurity of employment; alienation of youth; rejection of traditional values; and

252 Ibid., 238. 253 Jay Forrester, Urban Dynamics, 9, as quoted in Edwards, “The World in a Machine,” 238.

119 inflation and other monetary and economic disruptions.”254 Global in scope, deeply interrelated, and changing at an ever-accelerating rate, these issues constituted what Club member Hasan Ozbekhan defined as a “generalized meta-problem (or meta-system of problems) which we have called and shall continue to call the ‘problematique.’”255 The

Club thus came to call this collection of issues the monde problèmatique.

Limits to Growth was the group’s first and most famous attempt to provide a comprehensive analysis of this monde problèmatique. Funded by the Volkswagen

Foundation, the project took shape at a two-week conference in Cambridge,

Massachusetts in 1970. In Cambridge, Forrester presented the group with “World 1,” a rough computer model that described the world’s most important integrated systems.

Much as Forrester’s earlier urban and industrial models had, “World 1” demonstrated that unrestrained economic, agricultural, and demographic growth could not be sustained as a permanent systemic condition.256 Without rational management, the model showed, the world-scale boom that characterized the twentieth century would lead to a world-scale bust. Favorably impressed, the Club of Rome soon commissioned Forrester’s MIT

254 Donnella H. Meadows, Dennis L. Meadows, Jørgen Randers, and William W. Behrens III, The Limits to Growth: A Report for the Club of Rome’s Project on the Predicament of Mankind (New York: Universe Books, 1972), x. 255 As quoted in Elichirigoity, Planet Management, 76. 256 In “The World in a Machine,” Edwards describes how this idea of growth as a developmental stage rather than a constant developed in Forrester’s studies of industrial and especially urban systems. “Continued exponential growth” Edwards quotes from Forrester’s Urban Dynamics, “is impossible” (pg. 239). See also Elichirigoity, Planet Management, chapters 3, 5, and 6. It is also relevant to note that The Limits to Growth, while it was groundbreaking in its use of digital computers to simulate the coming catastrophe, was hardly alone in pointing to the problems of unrestrained growth. Perhaps most notably, biologists Paul Ehrlich and Barry Commoner engaged in a vituperative exchange during this time over the root cause of the coming environmental apocalypse. Commoner worried primarily about technology, Ehrlich and his colleague John Holdren worried about population. See Michael Egan, “When Scientists Disagree,” chapter 4 in Barry Commoner and the Science of Survival: The Remaking of American Environmentalism (Cambridge: MIT Press, 2007): 109-138. See also Paul R. Ehrlich and John P. Holdren, “Dispute,” Environment 14 (April 1972): 23-52; Roy Beck and Leon Kolankeiwicz, “The Environmental Movement’s Retreat From Advocating U.S. Population Stabilization (1970-1998): A First Draft of History,” Journal of Policy History 12 (2000), 123-156.

120 colleague and former student Dennis Meadows to build on Forrester’s world dynamics models in a computer model of his own designed to demonstrate the relationships between environmental degradation, population growth, and the impending failures of the world’s social, economic, and political systems.

As the name of the study suggests, the Limits to Growth group focused on the functional boundaries—or limits—of the world’s various interrelated large-scale systems.

Meadows’ group set out to examine the “five basic factors that determine, and therefore, ultimately limit, growth on this planet—population, agricultural production, natural resources, industrial production, and pollution.”257 The Limits to Growth group used a computer model very similar to Forrester’s updated world dynamics model, “World 2,” to demonstrate the inviability of continuous world-wide demographic, economic, industrial, and agricultural expansion.258 Meadows built off of the idea, promoted in the 1960s by

Eugene and Howard Odum, that all ecosystems trend toward equilibrium. The world, he argued, would soon be “faced with an inevitable transition from world-wide growth to global ecological equilibrium.”259 Without some sort of large-scale collective effort to curb growth intentionally, this transition could occur on its own—abruptly and catastrophically—sometime within the coming century. Donella Meadows, Dennis

257 Meadows et al., The Limits to Growth, xi. 258 Forrester soon updated “World 2” in another mode, “World 3,” which he described in detail in his more academic World Dynamics (Cambridge: Wright-Allen Press, 1971). In part, Forrester decided not to take the lead in The Limits to Growth in order to complete and publish World Dynamics with due speed. Unlike Limits, however, World Dynamics was more descriptive than proscriptive, and it sparkled with technical detail instead of burning with the political urgency of the Club of Rome study. Edwards, “The World in a Machine,” 243-245; Elichirigoity, Planet Management, 57. 259 Meadows made this argument explicitly in his initial proposal to the club of Rome (Quoted in Elichirigoity, 96), but the concept pervades the published Limits to Growth as well. The Odums, too, were systems scientists, and they both promoted the use of digital computers to numerically model the energy flow of ecosystems. Not coincidentally, Howard Odum served as a tropical meteorologist in the Air Force during WWII. He later credited his experience working with the large-scale, complex systems of the atmosphere for teaching him to think about other systems—particularly ecosystems—holistically. See Hagan, An Entangled Bank, 124-125.

121 Meadows’ wife, originally wrote the text of Limits to Growth as a sort of non-technical interpretive face sheet explaining the conclusions of the larger Club of Rome Study for the Volkswagen Foundation. Released a month before the Stockholm Conference, the published book sold more than seven million copies in thirty languages, and became a mainstay of popular environmentalism, both in the U.S. and abroad.260

The Limits to Growth described geographically discrete problems like soil depletion, mineral resource shortages, pesticide use, air and water pollution, overfishing, population growth, and nuclear waste in terms of a single, interdependent world system, but the study paid only cursory attention to the global atmosphere.261 Concurrent to their work on the Limits to Growth, however, a number of the Club of Rome’s members and consultants took a keen interest in the subject. In particular, Carroll Wilson of MIT’s

Sloan School of Management, a Club Member and former chairman of the Atomic

Energy Commission, recognized that atmospheric scientists’ previous success in using models to analyze and predict the behavior of non-linear feedback loops operating at local, regional, and global levels helped to validate the Club of Rome’s broader, holistic approach.262 What systems dynamicists sought to do with the Earth’s environmental, social, economic, and political systems more generally, atmospheric scientists had

260 Sales figures from Edwards, “The World in a Machine,” 244; McCormick, Reclaiming Paradise, 82. 261 The one reference to CO2 in The Limits to Growth is somewhat interesting, considering the theories available at the time. Models of the global atmosphere showed a probable relationship between CO2 and warming as early as the Manabe-Weatherald model of 1967 (discussed in ch. 1 and ch. 4). Limits to

Growth skirts the issue, however, noting that the atmospheric concentration of CO2 will likely reach 380ppm by the year 2000 (the actual amount was about 370ppm), but mentioning only that “this increase in atmospheric CO2 will eventually cease, one hopes before it has had any measurable ecological or climatological effect.” Limits clearly recognizes a potential effect, and dramatically displays an extended graph of exponentially increasing CO2, but uncharacteristically fails to make the consequences explicit. Meadows et al., Limits, 72-73. 262 Or as Edwards argues, “climate and weather models increasingly gave a picture of ‘the world’ as a whole, an interconnected set of systems whose interactions could be understood only through a combination of simulation and observation….In a certain epistemological sense, they gave us ‘the world’ as an ecological and physical unity.” “The World in a Machine,” 242.

122 already begun to accomplish with the Earth’s climate. Between 1970 and 1972, while

Limits to Growth was still in gestation, Wilson teamed up with atmospheric scientists, climatologists, and other management experts to incorporate the atmosphere—and by association, climate change—into a broader scientific vision of the threatened global environment.

The first and most important study to address climate as a major environmental issue was the 1970 Study of Critical Environmental Problems—the same SCEP study that addressed the atmospheric impacts of the SST. Principally organized by Wilson, SCEP was a veritable “who’s who” of atmospheric science, complemented by a smattering of notable management professionals (mostly from MIT), biologists, and scientific administrators from academia, government, and the private sector.263 Wilson, well- respected among scientists and well-connected from his days as the chairman of the AEC, brought this brain trust together for the month of July in Williamstown, , in order to “provide an important input” for the Stockholm Conference, which he and many of his colleagues found wanting in scientific objectives.264 In particular, Wilson and his eleven-man Steering Committee hoped to help “fill the gap” in research into “global problems such as changes in climate and in ocean and terrestrial ecosystems” that “had

263 In addition to Tom Malone and Roger Revelle, who are discussed below, the SCEP conference involved a good number of the nation’s most renown and influential atmospheric scientists, including Joseph Smagorinski of Princeton’s Geophysical Fluid Dynamics Laboratory, Hans Panofsky of Penn State, MIT’s Reginald Newell, Charles David Keeling of the Scripps Institution of Oceanography (famous for his measurements of atmospheric CO2), Wilmot Hess, Lester Machta, and J. Murray Mitchell of the Environmental Sciences Service Administration (later NOAA), Reid Bryson of the University of Wisconsin, Robert Fleagle of the University of Washington, G.D. Robinson of the Center for the Environment and Man, Inc., and many more. The conference also counted a number of powerful scientists and science advocates from other disciplines, like the biologist George Woodwell, Wilson himself (a former chief of the Atomic Energy Commission), and John L. Buckley of the President’s Office of Science and Technology, among its participants. SCEP, xvii-xxii. 264 SCEP, xi.

123 not been subjected to intensive study and examination.”265 The resulting book, published soon after the Williamstown gathering, synthesized more than 200 papers and scientific articles into a first assessment of “the status of the total global environment.”266

The SCEP group actually assessed two things. First, through working groups on the climatic and ecological effects of human activities, on the implications of environmental change and remedial actions, and on agricultural waste and energy production, the group outlined a rough schematic of what they saw as the major threats to

267 the global environment for the 1970s. The climatic effects of CO2, airborne particulates, and SST emissions figured prominently in this discussion, but so too did

DDT and other pesticides, mercury and other heavy metals, nuclear waste, more

“traditional” sources of air and water pollution like agriculture and industry, and, finally, oil spills.268 The SCEP report was in this sense a fairly comprehensive “state of the problem” assessment of the global environment—almost a sort of tentative, environment- specific preview of Limits to Growth.

But the SCEP authors were just as concerned with the effort to study the global environment as they were with the environment itself. In addition to their outline of the global environmental crisis, the SCEP authors provided a detailed assessment of the state of large-scale, systems-based environmental science. As Wilson explained in the preface to the published report,

“SCEP explored the procedures and programs of focused research, monitoring,

and action that will be required to understand further the nature of potential

265 Ibid. 266 Ibid. 267 Ibid., v-x. 268 Ibid.

124 threats to the global environment so that effective action can be taken to avert

future crises.”269

Steeped in the belief that better science would help make better policy, the SCEP group’s recommendations focused almost entirely on making more and better science.270 In general, they recommended a more extensive and cooperative international environmental monitoring effort, better data collection, and the development of rigorous, universal standards of observation and data analysis.271 Each working group also offered recommendations specific to research in their particular areas of expertise. For SCEP’s scientists, these programmatic recommendations were meant to set the parameters for understanding the global environment in the years to come.

Many of SCEP’s atmospheric scientists felt they had a natural role in helping to define the scientific parameters of the global environment. In a practical sense, atmospheric scientists studying weather and climate had already helped to create the physical scientific infrastructure necessary for measuring and monitoring certain basic elements of the environment on a global scale. In 1962, the United Nations asked the

World Meteorological Organization (WMO, a U.N. agency) to design and execute a large-scale, long-term program of global atmospheric research. Officially launched in

1968 as a joint venture of the WMO and the non-governmental International Council of

Scientific Unions (ICSU), the Global Atmospheric Research Program (GARP) employed satellites and computers to observe and model the Earth on a continuous and global

269 Ibid., xii. 270 For more on scientists’ belief in the power of scientific knowledge in policy-making, see Chapter 4. 271 For SCEP’s summary recommendations, see SCEP, 7.

125 basis.272 GARP also worked to build a terrestrial network of participating scientists who supplemented satellite data with on-the-ground observations from points all over the world.273 By providing a clearinghouse for global data, GARP established the channels of communication that connected the physical spaces of atmospheric research—weather stations, individual laboratories, and institutions—throughout the world. As a whole, this network of research facilities represented the material skeleton of a body of empirical knowledge necessary not only for studying the global atmosphere, but also for governing it.274 In the sense that scientists knew the Earth primarily through instruments, observations, and data, in the late 1960s this scientific skeleton essentially was the global environment.

GARP at once established the atmosphere as a global subject of study and developed a global network of scientists to study it, but for some ICSU leaders, GARP was only a first step. As early as 1968, two Americans in the ICSU were already beginning to think about ways to expand the GARP model of cooperative international atmospheric research in order to broaden this scientific image of the Earth’s environment.275 In 1968, Roger Revelle, then the Director of Harvard’s Center for

Population Studies, along with Tom Malone, a meteorologist, the Dean of the University

272 For more on GARP, see chapter 1. See also Erik Conway, ”Planetary Atmospheres,” in A History of Atmospheric Science at NASA, 1958-2004 (Baltimore, MD: Johns Hopkins University Press, 2008), 94- 121; Spencer Weart, The Discovery of Global Warming, (Cambridge: Harvard University Press, 2003), 99- 100; “ICSU and Climate Science,” on www.icsu.org; Thomas F. Malone, “Reflections on the Human Prospect,” http://humanprospect-post2.blogspot.com. 273 As Edwards is careful to point out, this data was not itself necessarily global—in fact, it was “inconsistent, poorly calibrated, and temporally brief”(246). Through their models, however, scientists made the “data function as ‘global’ by providing an overarching reference frame.” Local and regional data thus supported a global epistemology. 274 GARP ran for 15 years, and eventually led the U.N. to host a series of conferences on global climate change in the late 1970s—early predecessors to today’s IPCC. 275 GARP itself, however, remained focused on the atmosphere, and the project eventually evolved into a more climate-specific program, the World Climate Research Project (WCRP). Malone, “The Human Prospect.”

126 of Connecticut Graduate School, and one of the prime movers behind both GARP and later SCEP, lobbied the ICSU to begin incorporating global environmental concerns into their scientific programs.276 Malone and Revelle proposed a Scientific Committee on

Problems of the Environment (SCOPE)

“designed to cover environmental issues—either global or shared by several

nations—in urgent need of interdisciplinary syntheses through synthesis,

assessment, and evaluation of information available on natural and human-made

environmental changes and the effects of these changes on people.”277

Modeled after the Scientific Committees on Arctic and Ocean Research developed by the

ICSU in the 1960s, SCOPE would help coordinate international, interdisciplinary efforts to study and monitor the global environment in the 1970s.278 In 1969, with financial support from the Ford Foundation, the ICSU approved the new committee, and in 1970

Malone, then newly-elected as the Vice President of ICSU, was officially appointed

SCOPE’s secretary-general.279

276 Interview of Thomas F. Malone by Earl Droessler, February 18, 1989, AMS/UCAR Tape Recorded Interview Project. 277 Interview of Thomas F. Malone, February 18, 1989; ICSU website, http://www.icsu.org/5_abouticsu/STRUCT_InterBod_2.php?query=SCOPE 278 Global Environmental Monitoring: A Report Submitted to the United Nations Conference on the Human Environment, Stockholm 1972, Commission on Monitoring of the Scientific Committee on Critical Environmental Problems (SCOPE), International Council of Scientific Unions (ICSU) (Stockholm, Sweden: ICSU SCOPE, 1971). 279 Malone both promoted and embodied what many scientists saw as the ideal relationship between science and politics. An MIT-educated meteorologist trained in the milieu of the discipline’s war-time explosion in the 1940s, in the 1960s Malone served as a sort of ubiquitous, one-man interface between science and policy. By the time he took his position in SCOPE, he had already been President of both the American Meteorological Society (1960-61) and the American Geophysical Union (1962-64), Secretary-General of the ICSU’s Committee on Atmospheric Sciences (1964-68), Chairman of the National Science Foundation’s Advisory Panel on Atmospheric Sciences (1958-59), the Department of Transportation’s National Motor Vehicle Safety Council (1967-68), the U.S. National Committee on UNESCO (1965-67), a State of Connecticut Clean Water Task Force (1965-66), the Board of Directors of the Traveler’s Research Company (1958-1969) and the related Center for the Environment and Man (1969-71), and three National Academy of Sciences Committees. Now in his 90s, Malone still blogs about major scientific and environmental issues on his site, called The Human Prospect. A bio can be found at http://humanprospect- tfmbio.blogspot.com. See also Global Environmental Monitoring; www.icsu.org

127 Alongside SCEP, SCOPE helped to establish a place for global environmental monitoring and research next to the laundry list of more geographically specific issues like pesticides and waste management on the U.N.’s agenda. The ICSU had a strong history of close collaboration with the U.N., and by publicly identifying the myriad scientific concerns not yet addressed by Conference planners, SCEP helped to convince the U.N. that they needed the advice. In December of 1970, two months after the publication of the SCEP report, the Secretary-General of the U.N. Conference on the

Human Environment asked SCOPE to produce a “report recommending the design, the parameters and technical organization needed for a coherent global environmental monitoring system.”280 SCOPE’s first report, Global Environmental Monitoring, was tailored to that task.

Global Environmental Monitoring essentially offered a redux of the SCEP report, stripped down to only those findings and recommendations immediately relevant to the planning of the U.N. Conference on the Human Environment. Like SCEP, Global

Environmental Monitoring called for better systems of data collection, the establishment of international scientific standards, and better cooperation in environmental monitoring and research more generally. But SCOPE also outlined a specific plan for establishing a physical infrastructure of global environmental science under the auspices of the United

Nations with the help of its member states. The proposal included “an integrated network of reference stations” based on national monitoring efforts and centered around regional coordinating stations measuring, in addition to standard meteorological data, levels of atmospheric turbidity, solar radiation, mercury, lead, cadmium, DDT, and

280 Global Environmental Monitoring, 16.

128 polychlorinated biphenyls; the establishment of ten “baseline stations” for measuring future global environmental change; and two permanent “International Research

Reference Stations” to serve as home bases an training centers for the entire global research network.281 SCOPE also recommended programs to monitor factors of human health, including nutrition and fertility, as well as a comprehensive “biome studies” project designed to monitor everything from the distribution of endangered species to changes in land use and vegetation to bird migration patterns to microbial activity in local soils.282

Climate change quickly emerged as an important focus of the systems science approach. Of all the environmental problems they studied, the SCOPE group identified

“potentially adverse climatic change due to human activities” as the issue “most relevant”

283 to the project of international environmental monitoring. Atmospheric CO2 and aerosol content, the group advised, should constitute “first priority data” in even the most limited international environmental monitoring effort.284 SCEP, too, pointed to climate change as a “special case” in the list of global environmental problems, with “consequences for the

281 Malone originally envisioned these “International Research Reference Stations” as NCARs for the global environment. As he told Earl Droessler in a 1989 interview, Malone “hoped very much that we could get out of the Stockholm conference, say, an international research center of the UConn kind [it is likely he meant to say UCAR here], but at the international level.” Malone remembers that Senator Warren Magnuson supported his vision, and he, Malone, still hoped to build this kind of center when he moved from UConn to Butler University’s Holcom Research Center in 1973. Global Environmental Monitoring, 5-9; Interview with Thomas F. Malone, Feb 18, 1989; Telephone Interview of Thomas F. Malone with the author, June 4, 2009. 282 Global Environmental Monitoring, 10. 283 Global Environmental Monitoring, 22. 284 Interestingly, SCOPE’s recommendations for creating a monitoring infrastructure also centered around climate, although in a somewhat different way. Their report recommended that it’s ten “baseline stations” be partitioned based on climatic zones, including “(1) northern tundra, (2) northern coniferous forest, (3) northern hemisphere temperate grassland, (4) arctic or antarctic, (5) high mountain, (6) tropical forests, (7) desert or semi-desert, (8) tropical savanna or grassland, (9) oceanic island, (10) temperate deciduous forest.” Global Environmental Monitoring, 6-7.

129 human condition and human endeavor [that] could be enormous.”285 Again, the authors

identified CO2 and atmospheric aerosols as two of the most important new environmental problems for the 1970s.286 Despite the relative novelty of climate change, atmospheric scientists framed the issue as a top scientific priority.

Soon the problem engendered its own international study. Shortly after the

Williamstown meeting concluded, Wilson, Malone, NCAR’s William Kellogg, William

Matthews of M.I.T., and G.D. Robinson of the Center for the Environment and Man, Inc., launched a climate-specific follow-up to SCEP, the 1971 Study of Man’s Impact on

Climate (SMIC).287 Sponsored by M.I.T. and hosted by the Swedish Academy of

Sciences, SMIC endeavored to forge a consensus on “what we [scientists] know and do not know” about climate, and “how to fill the gaps.”288 The study’s report, published in a rush in 1971, offered specific suggestions on how to improve climatic and atmospheric research within the international scientific community, including the potential costs of its recommended research programs. With scientists representing 14 nations—including the

USSR—in attendance, SMIC had a more international character than SCEP, and because it dealt exclusively with climatic issues, it offered more detail in its scientific assessments. For the most part, SMIC mirrored its predecessor in its effort to provide

“input into planning for the 1972 U.N. Conference on the Human Environment,” and ultimately the study was a sort of second SCEP, focused specifically on climate.289 With

SMIC, however, scientists planted the climate and the atmosphere at the center of their definition of the global environment.

285 SCEP, 244-45. 286 Ibid., 4. 287 SMIC, xv. 288 Ibid. 289 Ibid.

130 The scientists involved in SCEP, SMIC, and SCOPE built their image of the global environment around the issue of climate change for a number of reasons. Most obviously, many of the prominent individuals who participated in the two studies were themselves atmospheric scientists or climatologists, and both the practical problems and professional ambitions of their disciplines guided their studies.290 As Stephen Schneider, the “rappateur” for SMIC, remembers in Science is a Contact Sport, promoting international efforts to study the global atmosphere served atmospheric scientists’ professional self-interest.291 Scientists studying climate needed a more advanced global

data network in order to validate and build upon the existing research into CO2 and atmospheric aerosols. “Atmospheric scientists had a common cause in all countries,”

Schneider recalls. “We needed expensive satellites, balloons, ships, and computers from our governments to do our work. International cooperation in data sharing reduced costs of these tools to individual nations.”292 When atmospheric scientists described the global environment in the early 1970s, it was no coincidence that what they envisioned matched their main research interests.

The SCEP and SMIC scientists’ focus on climate went beyond professional self- interest, however. Just as importantly, climate served as an intellectual anchor for the holistic, global-scale approach to environmental research increasingly advocated by

Malone, Wilson, and their colleagues. The concept appealed not only to atmospheric

290 The SST debate had made the climate—and CO2—a hot topic in the late 1960s, and Malone, Revelle, William Kellogg, and many others were already involved in studying the problem. Revelle, for example, while an oceanographer rather than an atmospheric scientists per se, had put CO2-induced climate change on the map with his 1957 Tellus article on the subject (see chapter 1). In 1970, Malone told a group at the California Institute of Technology, that, in the words of the L.A. Herald, “continued burning of fossil fuels will cause the Earth’s temperature to rise and create other grave climatic changes,” and that these changes might “threaten the human species.” Los Angeles Herald-Examiner, Monday, Oct. 19, 1970, A-13. 291 Stephen H. Schneider, Science as a Contact Sport: Inside the Battle to Save Earth’s Climate (Washington, D.C.: National Geographic, 2009): 3. 292 Ibid.

131 scientists, but to the studies’ biologists, oceanographers, and management professionals, too. Bent on providing input to the U.N. Conference on the Human Environment, SCEP sought out “environmental problems whose cumulative effects on ecological systems are so large and prevalent that they have worldwide significance.”293 The most “authentically global” of these environmental problems, the group found, was the problem of climate change.294 As SCEP pointed out, pesticide use, agricultural waste, and population growth were global problems prevalent in many different states, nations, and regions throughout the world; they were endemic, and thus in the aggregate, they were “global” problems.

But unlike climate, these problems had immediate causes and direct impacts, and they were typically rooted in discrete and independent geographical spaces.

The very concept of climate, by contrast—derived from the Greek, klima, or inclination, as in the inclination of the sun vis-à-vis points on the Earth—implied some supra-regional causal force beyond individual humans’ immediate perceptions.295

Climate change was not only a common problem, it was a problem determined by the

chemical composition of single common space: the atmosphere. CO2 and other chemical constituents circulated almost limitlessly in this ubiquitous mixture. As scientists like

Reid Bryson, Ishtiaque Rasool, Stephen Schneider, and others had by then begun to show, the potential impacts of climate change on the Earth’s natural and human systems were thus similarly limitless. Tied to the global atmosphere—fluid, dynamic, and borderless—climate change offered a prime example of the type of environmental problem that could only properly be addressed at a global scale.

293 SCEP, 5. 294 Ibid., 245. 295 See “climate, n.1” OED Online, Draft Revision, July 2010 (Oxford University Press, 2010).

132 Global Environmental Monitoring made the case bluntly. “A correct and complete evaluation of environmental data,” the authors argued, “is only possible when

296 the environment is treated as a unity.” CO2 and aerosols, recently identified in the SST debate as the most likely causes of global climate change, were themselves confined to the atmosphere, but their potential climatic impacts affected “all different media of air, water, soil, and biota including man.”297 As the international political community prepared to grapple with a complex and multifaceted global environmental crisis at

Stockholm, the SCOPE group argued that only an “integrated view” that combined research into all of the components of the Earth’s various large-scale systems could provide the appropriate information for understanding and managing threats to the global environment.298 The universality of climate underscored the need for such an integrated, systems-based global view.

The Politics of Systems Science

As an intellectual tool, the systems science concepts behind SCEP, SMIC, and

Global Environmental Monitoring served a mixture of scientific and political ends.

Through these studies, scientists advanced a detailed framework for studying the Earth’s threatened landscapes, natural resources, and species. 299 By setting the research parameters for measuring and monitoring the world’s large-scale systems, they made the global environment knowable. But each of these studies, written as preparatory documents for the 1972 U.N. Conference on the Human Environment, also engaged in an

296 Global Environmental Monitoring, 62. 297 Ibid. 298 Ibid. 299 SCEP and SMIC were intended as preparatory documents for the Conference; Limits to Growth was published, in something of a rush, shortly before the event, but not specifically for it.

133 ongoing political discussion about how the global environment—and by extension the world more broadly—ought to be governed.

For the SCEP, SMIC, and SCOPE authors, global environmental research and global environmental governance were two sides of the same coin. Not surprisingly, scientists favored the kind general international political cooperation that would facilitate their bold proposals for a program cooperative global environmental research. Leading systems scientists—the Odums in ecology, Forrester, Meadows, and Wilson in systems dynamics, and a growing cadre of scientists including Malone, Phil Thompson, William

Kellogg, and Steven Schneider in climate—shared in common a general commitment to using science as the basis for the rational management of agriculture, natural resources, and economic development across the globe.300 They also shared a common belief that the systems-science methodology was the best way to tackle these deeply interdependent problems. Strategies for controlling the world’s social, political, and environmental problems piecemeal or in isolation from each other, they contended, were bound to fail; planners must instead address local scale issues in terms of large-scale, interconnected systems.301 Large-scale management of interconnected global systems required more integrated, interdisciplinary research into these systems—that is, more systems science.

The political implications of the systems science approach went well beyond the drive for more and better cooperative research, however. Systems scientists presented the global environment in a way that all but demanded a broader paradigm of cooperative global governance. The borderless, interdependent global systems that scientists described contrasted sharply with the competitive, state-centered politics of the Cold

300 For more on climate scientists’ particular commitments to management, see chapter 4. 301 Edwards, “The World in a Machine,” 245.

134 War. On the contrary, systems science provided support for a movement within the international community to refocus and reinvigorate the United Nations in order to combat de facto superpower rule. Maurice Strong and many of his U.N. colleagues hoped that the U.N. Conference on the Human Environment would mark a transition away from the paradigm of divisive East-West politics toward a “new globalism” that revolved around a more robust and democratic system of international politics. When scientists like Meadows, Wilson, and Malone described the Earth as a deeply interdependent whole, they supported this global vision.302 In an interconnected world, actions by any one nation or even one industry affected the world’s predicament as a whole; conversely, the Earth’s systemic limits applied to all nations and groups in all circumstances.303 Borders, key points of control for the globalizing forces of trade, migration, and disease, had little meaning in the face of transnational environmental problems like air pollution, extra-national problems like ocean dumping and fisheries depletion, and truly global problems like ozone depletion and climate change. Unilateral decisions, belligerent nationalism, and Cold War posturing had no place in a finite world

302 Conference Secretary Maurice Strong and many of his U.N. colleagues welcomed systems scientists’ holistic message, and they made an effort to incorporate systems scientists’ organizational recommendations into their larger agenda for Stockholm. In September of 1971, the U.N. invited Malone, Wilson, and SCOPE’s Bengt Lundholm to participate in a “Conference on International Organization and the Human Environment,” attended by, among others, U.N. General Secretary U Thant. Malone and Lundholm attended the conference; Wilson did not. Held in New York, the meeting helped to reinforce the importance of international cooperation and collective action in tackling the world’s global environmental problems. Not coincidentally, the discussion outline for the conference, penned by Richard Gardner of Columbia University, identified atmospheric pollution as a “nuisance of broad international significance” in its section on “priorities for action and institutional implications.” “List of Persons Invited to the International Organization and the Human Environment,” Items-in-lnternational Organization on the Human Environment's Conference opening, 20 May 1971, Operational Files of the Secretary-General: U Thant: Speeches, Messages, Statements, and Addresses - not issued as press releases; “International Organization and the Human Environment,” Items-in-lnternational Organization on the Human Environment's Conference opening, 20 May 1971, Operational Files of the Secretary-General: U Thant: Speeches, Messages, Statements, and Addresses - not issued as press releases. 303 The SMIC authors hung their hats on the idea of rationality, and they believed that international cooperation represented the only “rational” response to “a set of decisions that could govern the future habitability of our planet.” If they did not believe that society would be rational, the group admitted, “The exercise we have begun would be fruitless indeed.” SMIC, 27.

135 composed of deeply integrated social, political, and environmental consequences. SCEP,

SMIC, and Limits to Growth thus backed up the “Only One Earth” slogan of the U.N.

Conference on the Human Environment with a “we’re all in this together” form of science.304

Maurice Strong, the U.N., and “a new kind of globalism”

Maurice Strong was more explicit about his political goals for the U.N.

Conference on the Human Environment than were the scientists that tacitly supported him. Though Strong himself consistently demonstrated a central concern for scientific research in the service of environmental protection, the “global vision” that shaped his approach to Stockholm reflected a set of geopolitical values that had as much to do with the Cold War as they did with the growing and poorly defined litany of pandemic environmental problems that constituted the so-called “global environmental crisis.”305

Strong and his colleagues saw in the pressing but poorly defined environmental crisis the type of broad, collective, international problem that the now ailing U.N. had originally been designed to address. He hoped to use the Stockholm Conference as way to re-

304 The Limits to Growth made this connection explicit. The published work began with former U.N. Secretary U Thant’s call for “a global partnership to curb the arms race, to improve the human environment, to defuse the population explosion, and to supply the required momentum to development efforts.” It ended with a call for the international community to create a cooperative “world forum where statesmen, policy-makers, and scientists can discuss the dangers and hopes for the future global system without the constraints of formal intergovernmental negotiations.” Meadows et al., Limits, 21, 200; Only One Earth: The Care and Maintenance of a Small Planet (New York: W.W. Norton & Co, 1972) was the name of a book by Rene Dubos and Barbara Ward, written specifically for the 1972 U.N. Conference on the Human Environment. The phrase was also the slogan for that meeting. 305 Strong was primarily concerned with international politics, but he earned a glowing reputation among scientific leaders as well. In a recent telephone interview with the author, Tom Malone recalled with deep reverence Strong’s ability to move back and forth between the scientific and political communities with ease. “He deserves a gold medal for what he did,” Malone said. “He was man who had a profound impact on civilization in my opinion.”

136 empower the United Nations as a force for international peace and cooperation by mobilizing its institutional machinery to deal with the world’s environmental problems.306

Strong took the job of Conference Secretary in January of 1971 as a rising star within the organization. That same month, the man who appointed him, U Thant, announced that he would not serve a third term as U.N. General Secretary, and Strong became one of the leading candidates for the job. Strong was a rags-to-riches businessman whose “disarming modesty and beguiling punctiliousness,” according to

The New York Times, belied a “phenomenal dynamism and an extraordinary skill at both business and diplomacy.”307 He cut an unimposing figure—slightly built with receding dark hair, thick, close-cut sideburns and a toothbrush mustache—but his presence made an impression. He was a professor at York University in Toronto, though he never went to college himself.308 He spoke Inuit.309 A former president of one of his native Canada’s biggest utilities companies, the 42 year-old’s extensive business experience and his myriad contacts in the oil and gas industries made him an appealing candidate to U.N. members concerned about international development. His success as the head of

Canada’s international aid program—which grew from $80 million to $400 million under his directorship—complemented his business resumé.310 Strong’s initial success in winning support for the Conference among U.N. members, combined with his delicate handling of tensions that arose between the developed and developing world at a meeting in Founex, Switzerland (discussed below), earned him a reputation as an honest and

306 Gladwin Hill, “U.S. to Suggest Pollution Board Reform,” The New York Times, Dec. 8, 1970. 307 “Planner of Global Talks: Maurice Frederick Strong,” The New York Times, Sept. 23, 1971. 308 Ibid. 309 Ibid. 310 Henry Tanner, “Canadian Added to Candidate List for Thant Job,” The New York Times, Nov. 19, 1971.

137 dynamic player in international politics.311 Though Kurt Waldheim of would eventually beat out Strong to succeed Thant, Strong articulated a vision for the future of the organization shared by many in the General Assembly.312

Strong envisioned the U.N. Conference on the Human environment as the centerpiece of a larger effort to promote the goals of global international cooperation, democracy, and unity within the hierarchy of U.N. values. Central to this effort was the strengthening of the General Assembly, where the U.N. dealt with cultural, social, and now environmental concerns. Real power within the U.N. was divided unequally between three main groups: the five permanent members of the U.N. Security Council

(France, the United Kingdom, China, the United States, and the Soviet Union); a set of international financial institutions responsible for financing development and maintaining stable monetary exchange rates called “Bretton Woods Institutions;” and the all-inclusive

General Assembly.313 Despite its size, the General Assembly was the least powerful of

311 “Planner of Global Talks: Maurice Frederick Strong,” The New York Times, September 23, 1971; Maria Ivanova, “Looking Forward by Looking Back: Learning from UNEP’s History,” in Global Environmental Governance: Perspectives on the Current Decade, edited by Lydia Swart and Estelle Perry (New York: Center for UN Reform Education, 2007), 26-47. 312 Waldheim’s election as the Secretary-General was unsurprising, but, in the end, quite controversial. Unlike Strong, Waldheim didn’t intend to “make waves” at the U.N., and he was an attractive choice for members of the powerful Security Council, which included the U.S., China, and the Soviet Union. But Waldheim refused to talk about his past, and only later was it revealed that he had served as an intelligence officer in the Wehrmacht during the Second World War, and that he had served in units shown to be involved in atrocities in the Balkans. Waldheim’s role in Balkan War crimes was never sufficiently proven, but he was shown to have lied about his war record. These issues arose during his second campaign for President of Austria in 1985, in what became known as “The Waldheim Affair.” He was nevertheless successful in his campaign, and served as Austria’s President until 1992. Paul Gordon Lauren, “Diplomats and Diplomacy of the U.N.,” in The Diplomats, 1939-1979, edited by Gordon A. Craig and Francis L. Lowenheim, Princeton: Princeton University Press, 1994, pg. 459-495. For more on the “Waldheim Affair,” see E.M. Rosenbaum with W. Hoffer, Betrayal: The Untold Story of the Kurt Waldheim Investigation and Cover-Up (New York: St. Martin's Press, 1993). 313 The most powerful of these bodies, the Security Council, counted the Soviet Union, the United States, China, France, and the United Kingdom as permanent members. The Security Council maintained preeminence in matters involving military activities and “peacekeeping,” and its affairs had dominated the internal politics of the United Nations since its inception in 1945. Technically under the purview of ECOSOC and not an independent U.N. organ, the Bretton Woods Institutions—mostly banks—ensured that economic instability could not upset the power relationships within the Security Council. These institutions

138 the three. Like many of his colleagues from “third-party” countries in Scandinavia,

South America, and parts of Asia, Strong objected to the U.N. Security Council’s dominance of U.N. politics. Nations with no nuclear weapons and no representatives on the Security Council criticized this small governing body as undemocratic. They unfavorably contrasted it with the “one nation, one vote” system of the General

Assembly. They feared that the superpowers’ central concern over “peacekeeping” and international security undermined the other functions of the United Nations as a whole.

Strong called for a consolidation of the U.N.’s resources in order to better address the non-military policy areas that concerned the General Assembly; he hoped that the

Conference on the Human Environment would at once help revitalize this “soft side” of the U.N.’s mission and give the nations of the General Assembly a stronger voice in the organization’s long-term goals.314

Strong presented the U.N. Conference on the Human Environment as a transformative moment in international relations. “The principles of the U.N. charter have remained largely unfulfilled,” Strong told The New York Times in 1970, “because

also oversaw economic development. The Bretton Woods Institutions had been set up essentially to allow the U.S. to provide easy credit to the capitalist governments of developing nations. At the moment of the Conference in they were in the process of unraveling after Nixon refused to sell gold on the international market at the price stipulated by the IMF in 1971. Ironically, it would be to the descendents of these institutions that both governments and non-governmental environmental organizations would begin to turn to for help on environmental issues during the conservative backlash of the 1980s (more on this in chapters 5 and 6) Odd Arne Westad, The Global Cold War: Third World Interventions and the Making of Our Time (New York: Cambridge University Press, 2005), 155-6. For more on Bretton Woods institutions and on the history of the U.N. more generally, see Paul Kennedy, The Parliament of Man: The Past, Present, and Future of the United Nations, New York: Random House, 2006; Stanley Meisler, United Nations: The First Fifty Years (New York: The Atlantic Monthly Press, 1995). 314 Paul Kennedy uses the environment as a prime example of what he calls “the soft face of the U.N.’s mission.” In retrospect, he argues, the U.N.’s emphasis on fostering social and cultural development may seem naïve, but in the context of rebuilding Europe in 1946, these goals made a lot of sense. Once the Cold War became the central issue of U.N. diplomacy, however, these goals faded, Their revitalization in the 1970s was complicated by demands from the less developed world that organizations like the WHO, UNICEF, and UNESCO account for their concerns rather than solely those of the developed world. Kennedy, Parliament of Man, 143-176; Henry Tanner, “Canadian Added to Candidate List for Thant Job,” The New York Times, Nov. 19, 1971.

139 men and nations perceived their greatest threats to lie in other men and nations.”315 The crisis of the global environment—a crisis made up of the myriad and interrelated threats to the air, water, land, and people of nations around the world—offered the United

Nations an opportunity to transcend East-West political tensions by working in concert toward the common objective of environmental protection. “If the member governments of the United Nations can find the political will to do so,” Strong proposed, “they can make the conference on the human environment the starting point of a world mobilization for the future of man.”316 Cooperation, Strong argued, lay at the very foundation of the

United Nations’ mission. And so too, he suggested (echoing the systems science creed), did cooperation underpin a healthy environment. “The entire global system on which all life depends,” he pronounced in 1971, “must inevitably and inexorably lead us back to a new kind of globalism.”317

UNCHUE, Development, and the Global Environmental Crisis

Strong’s “new globalism”—a vision of cooperative international politics led by a reinvigorated, more democratic United Nations focused, at least a first, on environmental issues—meshed almost seamlessly with systems scientists’ holistic image a fragile, solitary Earth made up of complex and deeply interconnected environmental systems.

The U.N. Conference on the Human Environment provided scientists with a focal point for their arguments in favor of a more cooperative and integrated approach to global

315 Strong pointed specifically to Article IX and Article X of the U.N. Charter. Article IX introduces the organizational objectives of economic and social cooperation, while Article X defines the structure and function of the Economic and Social Council. A “Charter of the United Nations” can be found as an appendix in Kennedy, The Parliament of Man, 313-341; Hill, “Pollution Board Reform.” 316 Hill, “Pollution Board Reform.” 317 Gladwin Hill, “U.N. Parley Plans a Global Environment Conference,” The New York Times, Sept. 14, 1971, pg. 5.

140 environmental research. Their recommendations, in turn, supported U.N. leaders’ efforts to foster a more cooperative and democratic approach to global environmental governance by underscoring the futility of piecemeal, national-level responses to environmental problems. As the Conference approached, however, this new global vision ran headlong into the realities of an international political system built upon regional and national self-interest. Within this competitive system, the rubric for what constituted an important environmental issue was hotly contested. U.N. member states sought to promulgate their national economic and political objectives by recasting the key problems of this new global space in terms of their pre-existing political concerns.

Stockholm thus became as much a competitive struggle to define the global environment as it was a cooperative effort to protect it.318 Not surprisingly, the most important issue in this contest involved the problems of economic development.

The original inspiration for the U.N. Conference on the Human Environment arose out of a Swedish challenge to nuclear development in the United Nations General

318 In its simplest and most literal definition, the term “global environment” is almost meaninglessly vague. Etymologically, it would be difficult to conceive of two more general words. From the French “environ,” or “to form a ring around, surround, encircle,” the “environment” is in its simplest and oldest definition “that which environs; the objects or the region surrounding anything.” “Global,” meanwhile, from the Latin globus, is in its most literal sense an adjective describing a round mass, ball, or sphere, although it has also come, after the French usage, to describe a thing or set of things in their comprehensive, all-encompassing, or unified totality. More recently, beginning in the late 19th century, the word began to pertain more specifically to things whole, world-wide, or universal. So in a sense, the “global environment” is nothing less than the unified and universal totality of all those things that influence…well, everything. The term “global environment” does, however, carry a very specific meaning in at least one disciplinary lexicon—that of computer programming. In programming, a “global environment mechanism” or “global environment paradigm” is a situation in which a single variable is accessible in every scope—that is, any part of a program can affect the variable, and the variable can affect any part of the program. The definition, not surprisingly, resonates with the “global dynamics” concept of systems science, where relationships are highly complex, often non-linear, and most importantly deeply interrelated. See “global, adj,” OED Online, June, 2010, Oxford University Press, 17 July, 2010, http://dictionary.oed.com/cgi/entry/50095613; “environment,” The Oxford English Dictionary, 2nd Edition, 1989, OED Online, Oxford University Press, 17 July, 2010, http://dictionary.oed.com/cgi/entry/50076498.

141 Assembly in 1967.319 Inga Thorsson, a Swedish diplomat, and Sverker Åström, Sweden’s permanent representative to the U.N., objected to an assembly proposal for a conference on the peaceful use of atomic energy, on the grounds that it would benefit only a limited number of nations who had nuclear industries. The Swedish delegation suggested an alternative conference that would “focus the interest of member countries on the extremely complex problems related to the human environment.”320 The Swedish government, apprised of the situation after the fact, launched an initiative on the issue within ECOSOC in 1968. They proposed to hold such a conference in 1972.321 Sweden offered to host the conference in their capital city, and to provide a significant financial contribution to its execution. ECOSOC, in turn, submitted the proposal to the General

Assembly, which unanimously approved the measure on April 2, 1969.322

As its title suggests, the focus of the U.N. Conference on the Human Environment was unabashedly anthropocentric. Moreover, scientists’ efforts to study the Earth’s truly global spaces notwithstanding, the environment that U.N. member states worried about consisted mostly of a compendium of local and regional issues—issues that they believed represented, in the aggregate, a human crisis of global proportions.

Problems of the human environment were many and varied, however, and prioritizing these problems turned out to be no easy task. Initial support for the proposal in ECOSOC reflected a growing international interest in the middle-class quality of life

319 In Reclaiming Paradise (47-48), John McCormick argues convincingly that the idea of a threatened global environment grew in part out of an international milieu of fear—and in particular, out of the Cold War fear of nuclear annihilation. The nuclear genesis of the U.N. Conference on the Human Environment bears McCormick’s point out, and it is no coincidence that the debate about nuclear disarmament played an important role in defining the global environment at Stockholm. 320 Ivanova, “Looking Forward,” 3, quoting Åström. 321 Ivanova, “Looking Forward,” 3. “U.N. Group Urges Curb on Pollution,” The New York Times, July 31, 1968; “Thant Urges Concerted Action on Pollution Crisis,” The New York Times, June 24, 1969. 322 “The Proceedings in the U.N.,” The New York Times, April 2, 1969.

142 issues that also lay at the core of mainstream American environmentalism.323 Europeans watched as their rivers caught fire and their forests died.324 Industry polluted their air and their water.325 In concert with growing political support for international cooperation and nuclear disarmament, the “progressive” nations of Europe began to value humans’ interactions with the natural world as part of a larger picture of a healthy society.

“Environmental” problems were at once natural and social. Åström pointed to uncontrolled urban growth as one of Sweden’s most pressing environmental problems.

He discussed air and noise pollution alongside increased traffic congestion and increased accident rates. Environmental degradation, Åström argued, also caused “problems connected with family disorganization, mental tensions, and increased crime rates.”326

U.N. Secretary General U Thant worried that an exponentially increasing population would accelerate this environmental degradation, stressing ecosystems, natural resources, and, most importantly, the stability of modern societies.327 With so many industrial nations in close proximity, environmental issues that remained largely regional problems

323 For more on the driving forces behind American environmentalism in the 1960s and ‘70s, see Samuel P. Hays, Beauty, Health, and Permanence: Environmental Politics in the United States, 1955-1985 (Cambridge University Press, 1987); Robert Gottlieb, Forcing the Spring: The Transformation of the American Environmental Movement (Washington, D.C.: Island Press, 1993); Kirkpatrick Sale, The Green Revolution: The American Environmental Movement, 1962-1992 (New York: Hill & Wang, 1993); and Roderick , Wilderness and the American Mind (4th Edition) (New Haven: Yale University Press, 2001). 324 Nash discusses some of these issues in chapter 16 of Wilderness and the American Mind, entitled “The International Perspective,” 342-379. Nash focuses primarily on preservation and land use issues, however, especially in Africa, rather than on Europe per se. As Douglas Weiner shows excellently in his study of Russian environmental politics, A Little Corner of Freedom: Russian Nature Protection from Stalin to Gorbachev (Berkeley: University of California Press, 1999), environmental activism in Europe was often just as tied to other political movements there as American environmentalism was tied to opposition to the war and the concern over free speech and civil liberties in the U.S. See Christof Mauch, Nathan Stoltzfus, and Douglas R. Weiner (eds), Shades of Green: Environmental Activism Around the Globe (Lanham, MD: Rowman & Littlefield Publishers, 2006). See also Ramachandra Guha, Environmentalism: A Global History (New York: Longman Publishing Group, 1999). 325 Ivanova, “Looking Forward,” 27. 326 “U.N. Group Urges Curb on Pollution,” The New York Times, 7/31/68. 327 “Thant Urges Concerted Action,” The New York Times 6/24/69.

143 in the United States required international regulatory mechanisms in Europe. Many believed the United Nations was the right organization for the job.

The Conference found wide support in principle, but divisions quickly arose over what the developing world perceived as a First World bias. Questions arose about the specific definition, causes, and appropriate U.N. responses to environmental degradation.

Disagreements revolved around the relative priorities of environmental protection and economic development. In March of 1970, a 27-nation Conference preparatory committee met in New York to select particular topics for the Conference and to decide upon its organizational structure.328 Though over half the nations represented in the committee came from the developing world, the program primarily reflected the interests of the United States, the Soviet Union, and the nations of Western Europe. Less developed countries objected. LDCs argued that the United States, the Soviet Union, and

Europe bore the primary responsibility for problems like industrial pollution, uncontrolled urban development, and toxic waste management. These issues also disproportionately affected First World nations. In short, these were the problems of economic success.

Developing nations, which made up a majority of the United Nations’ member states, had a set of environmental problems of their own. These were equally anthropocentric concerns, but they were the problems of poverty. Unlike the “quality of life” issues facing the developed world, the LDCs’ concerns over poor water, poor

328 The preparatory committee consisted of representatives from Argentina, Brazil, Canada, Costa Rica, Cyprus, Czechoslovakia, France, Ghana, Guinea, India, Iran, Italy, Jamaica, Japan, Mauritius, Mexico, the Netherlands, Nigeria, Singapore, Sweden, Togo, the Union of Soviet Socialist Republics, the United Arab Republic,' the United Kingdom of Great Britain and Northern Ireland, the United States of America, Yugoslavia, and Zambia. “Constitution of the Conference,” UNEP website http://www.unep.org/Documents.Multilingual/Default.asp?DocumentID=97&ArticleID=1496&l=en

144 sanitation, poor nutrition, and poor public health involved the maintenance of life itself.329

In contrast to First World “amenity” issues that stemmed from unchecked economic growth, they argued, economic gains would actually alleviate these Third World environmental stresses. LDCs worried that international environmental regulations would exacerbate these particular environmental problems by upsetting trade and stifling development. Many advocated a Second United Nations Development Decade in the

General Assembly to pick up in the 1970s where the First Development Decade, the

1960s, had left off.

Some LDCs, especially in Africa, had gained political independence as recently as the 1960s, and they complained that the First World was using “the environment” to deny them access to economic advancement and political security.330 “Environmental concerns,” critics lamented, “were a neat excuse for industrialized nations to pull the ladder up behind them.”331 LDCs had little interest in participating in a conference dealing exclusively with First World concerns about things like urban growth, recreational open spaces, endangered species, or industrial air pollution; they were even less keen if the resulting environmental regulations might stand in the way of their economic development.

The First and Third Worlds found common ground in a mutual concern over the exploitation of natural resources like timber, oil, and various minerals, but even here, there was a significant divide between the priorities of developed and developing nations.

While the LDCs worried that environmental regulations would impede development,

329 Development and Environment: Report and Working Papers of a Panel of Experts Convened by the Secretary-General of the United Nations Conference on the Human Environment (Founex, Switzerland, June 4-12, 1971) (Paris: Mouton, 1972): 1, 4. Hereafter simply “The Founex Report.” 330 Ivanova, “Looking Forward,” 29. 331 Ibid., quoting Åström.

145 many conservationists in the industrial world feared that the economic interests of development would overshadow environmental protection, especially within the development-friendly context of the United Nations. In addition to its primary roles of peacekeeping and diplomatic power-broking, during the “United Nations Development

Decade” of the 1960s the U.N. had essentially served as an international development organization. Now, critics like George F. Kennan questioned the compatibility of this focus on development with the goals of international conservation. “There is a considerable body of opinion, particularly in U.N. circles,” Kennan explained in a 1970s

Foreign Affairs essay, “To Prevent a World Wasteland,” “to the effect that it is a mistake to separate the function of conservation and protection of natural resources from that of the development and exploitation of these resources for productive purposes.”332 Kennan, the dean emeritus of American foreign policy, disagreed. “This is an area,” he argued,

“in which exploitative motives cannot usefully be mingled with conservational

ones. What is needed here is a watchdog; and the conscience and sense of duty of

the watchdog must not be confused by contrary duties and undertakings. It may

be boldly asserted that of the two purposes in question, conservation should come

first. The principle should be that one exploits what a careful regard for the needs

of conservation leaves to be exploited, not that one conserves what a liberal

indulgence of the impulse to development leaves to be conserved.”333

Nations, Kennan continued, had a disincentive to honestly and thoroughly study and regulate environmental deterioration, especially those developing nations with the most to lose economically through regulation. With stable economies and scientific resources

332 George F. Kennan, “To Prevent a World Wasteland: A Proposal,” Foreign Affairs, vol. 48, no. 3 (April, 1970), 407-408. 333 Ibid.

146 at their disposal, Kennan concluded, a relatively small group of advanced industrial nations—whose economies, after all, produced the bulk of the environmental problems—should take the initiative to study and correct environmental problems outside of the pro-development framework of the United Nations.334

Kennan’s focus on the conservation of natural resources might have been more in line with the interests of the Third World than Åström’s concerns about traffic and urban growth, but there was still an important difference in perspective. Like most of his First

World colleagues, Kennan still saw the global environmental crisis primarily as a consequence of more than a century of unchecked industrial development. LDCs, who had yet to experience industrialization, focused instead on the environmental consequences of underdevelopment. And it was the LDCs who, for once, had the upper hand.

Unlike Kennan, many First World leaders had by 1970 already committed to using the machinery of the United Nations to create a framework for international environmental regulation, and these leaders realized that for such an international effort to be successful, they needed the participation of the developing world. LDCs made up a majority of the United Nations’ membership, and produced many of the material resources potentially subject to regulation. In a democratic United Nations, the human environment that U.N. leaders like Strong and U Thant sought to outline would have to reflect visions of the environment from both sides of the industrial divide. In June of

1971, a “Panel of Experts on Development and Environment” met at Founex,

Switzerland, to “consider the protection and improvement of the environment in the

334 Ibid., 408.

147 context of the urgent need of the developing countries for development.”335 Worried that some LDCs might choose not to participate after a contentious meeting of the Conference preparatory committee the previous February, the informal panel sought to reassure the developing world that their interests would be represented in Stockholm. The meeting resulted in one of the foundational documents of contemporary international environmental politics, the “Founex Report.”

The Founex Report articulated three tiers of environmental concern, all subservient to development.336 The highest priority for developing nations was to ameliorate those basic environmental problems associated with rural and urban poverty that could, in large part, be solved by development itself. These included poor sanitation, disease, malnutrition, and poor water quality. The panel also recognized that the process of development had environmental implications, however. Each step of development—agricultural growth, industrialization, and eventually integrated transportation and communication networks—put strains on developing nations’ natural and human environments, and these problems constituted the second tier of developing world environmental concerns. “Resource depletion,” “biological pollution,” “chemical pollution,” “physical disruption,” and “social disruption,” the report argued, represented potential threats to the very societies the LDCs hoped to build through development.

These problems would have to be weighed against nations’ continuing economic gains.337

The environmental concerns of industrially advanced nations like the United Sates, meanwhile, were not even identified, but instead tersely dismissed as “amenity” issues of

335 “Constitution of the Conference” 336 Development and Environment: Report and Working Papers of a Panel of Experts Convened by the Secretary-General of the United Nations Conference on the Human Environment (Founex, Switzerland, June 4-12, 1971), (Paris: Mouton, 1972), 1.2. Herafter, the “Founex Report.” 337 Founex Report, 2.6

148 a lower order. “The major environmental problems of developing countries,” the report stated flatly, “are essentially of a different kind.”338

In one sense, Kennan saw his concerns about the dominance of development manifested in the Founex Report. At Founex, the LDCs used the threat of non- participation as a way to ensure that international environmental protection served the interests of economic development. The report made it clear that Third World participation required substantial First World investment in developing economies. They appealed for a more “equitable” sharing of the increased costs of environmentally sensitive development, and suggested that increased development assistance would help ease the burden of environmental protection. As Kennan feared, the U.N. Conference on the Human Environment looked like little more than a sideshow in what would be a second U.N. Development Decade.

But the Founex Report also articulated a deep concern among the LDCs over a certain set of environmental issues that Kennan had not included in his analysis for

Foreign Affairs. In part, Kennan’s argument for an international environmental organization outside of the United Nations hinged on the developing world’s lack of interest in environmental protection. As the Founex Report revealed, that was simply not the case. LDCs continued to give economic development the highest priority, but they also expressed urgent concerns over the environmental problems associated with poverty and underdevelopment, as well as over those related to the process of development itself.

“The problems are already severe enough in developing countries,” the Founex Report read,

338 Founex Report, 1.2

149 “but in the absence of resolute action, they will tend to attain formidable

dimensions in the decades ahead…They can only aggravate the serious social and

political tensions that even now prevail in these societies. There can indeed be little

doubt about the urgent need for corrective action.”339

Recast in terms of basic needs like safe drinking water, improved sanitation, better nutrition, and a more efficient use of land and natural resources—needs that fell in line with the larger social objectives of economic development—the health of the environment became a pressing and important economic issue for the LDCs.

The issues outlined in the Founex Report were strikingly similar to the main problems then being identified by the Club of Rome in the Limits to Growth study, and the Founex Report’s authors highlighted the same concept of interdependence that stood at the core of the systems science approach. Agricultural and industrial development exacted a great burden on natural resources, but as the LDCs’ demands demonstrated, the environmental systems they damaged could not be divorced from the economic systems they supported. The interdependence of systems, they showed, ran both ways. The report also reflected a willingness to consider the environment alongside social and political conditions as an important non-economic factor in development. “Whilst the concern over the human environment can only reinforce the commitment to development,” the Panel of Experts contended, “it should serve, however, to provide new dimensions to the development concept itself.”340

Despite its similarities to The Limits to Growth, however, the Founex Report represented an unequivocal validation neither of Strong’s cooperative global vision nor of

339 Founex Report, 1.9 340 Founex Report, 1.6

150 the of the implicit politics of systems science that supported it. Rather, the meeting at

Founex stood out as a reminder that any regime of global environmental protection required global political buy-in, and that buy-in hinged upon the traditional national and regional interests of individual member states. The term “global” referred as much to the geopolitical interests at stake as it did to any kind of single or unified geographical space.

Systems scientists presented a world with environmental problems that could only be addressed on a global scale, but their internationalist ideals were untenable in the face of a political reality in which there was little agreement on what constituted a high priority environmental issue in the first place. The developing world was ready to commit to a common global effort to protect the environment, but only if “the environment” could be made meaningful to a developing world political constituency overwhelmingly concerned with issues of national economic and social advancement.341 The very self-interest that

Kennan saw as a threat to goals of international conservation efforts could, if properly managed, help solidify support for a geopolitically global regime of environmental protection. In “To Prevent a World Wasteland,” Kennan argued that only “the great international media of human activity”—the oceans and the atmosphere—might effectively be governed by an international organization like the United Nations.342 But truly geographically global problems like atmospheric and climatic change, “subject to the sovereign authority of no national government,” had no interested political constituency. Not surprisingly, global climate change found no place in the Founex

Report—nor in Kennan’s public about-face on the value of the U.N. Conference in

March of 1972.

341 “Kennan Now Backs An Agency in U.N. on World Pollution,” The New York Times, March 3. 342 Kennan, “To Prevent a World Wasteland,” 405.

151 The U.N. Conference on the Human Environment, June 5-June 16, 1972

The U.N. Conference on the Human Environment was the international community’s first attempt to deal with the environment as meaningful global political issue, and the framework for environmental governance established at Stockholm continued to guide discussions about the global environment for nearly four decades.

When it finally got underway after two years of careful preparation in June of 1972, the event involved two types of discussions about the environment. The first set of discussions revolved around the establishment of practical international mechanisms to monitor and control environmental degradation, and here, the U.S. took the lead.

Emphasizing institutional coordination, cooperative scientific research, and a few key environmental issues that resonated with Nixon’s domestic constituents, the U.S. delegation introduced a set of six limited but pragmatic initiatives for building an international environmental bureaucracy much like the one that Nixon had established at home.343

Despite a dearth of professional scientists in the delegation’s ranks, the

Administration’s proposals included significant support for international environmental science—and, in particular, for research into the global atmosphere. In Stockholm,

Russell Train and his colleagues unveiled “Earthwatch,” an international program of scientific cooperation intended to provide a framework for assessing environmental problems on a global scale. Earthwatch had two main objectives. The first was

343 White House Deputy Assistant John C. Whitaker outlined these initiatives in a memo to the President via John Ehrlichman shortly after the conference ended. The memo declares the U.N. Conference a success for the U.S. delegation, and in fact, the goals outlined in Whitaker’s memo do reflect the goals outlined in the delegation’s “Scope Paper” heading into the Conference. Memorandum for the President from John C. Whitaker via John D. Ehrlichman, June 23, 1972, Nixon Presidential Material Project, White House Central Files, Staff Member Office Files, John Whitaker Papers; “United Nations Conference on the Human Environment, Stockholm, June 5-16, 1972, Scope Paper,” John C. Whitaker Papers.

152 essentially bureaucratic. Much as the National Oceanic and Atmospheric

Association—created by Nixon in 1970—had streamlined redundant state and federal research efforts at home, the U.S. proposed that Earthwatch coordinate, expand, and reorganize existing national and international environmental research efforts at the United

Nations. The second objective of Earthwatch was programmatic, and it involved launching new, collaborative efforts to “measure trends and identify problems requiring international action.”344 In particular, Earthwatch focused on the potential threats to the oceans and the atmosphere identified as key issues in the SCEP and SMIC studies.

Robert White, director of NOAA and a member of the U.S. delegation, identified the

accumulation of dust particles, a rise in atmospheric CO2, and a decline in atmospheric ozone as three of the program’s primary concerns.345 The World Meteorological

Organization, a specialized U.N. agency, would take the lead in this new research.

On the surface, Earthwatch looked much like the type of large-scale program of global environmental research that systems scientists had called for in SMIC, SCEP, and

Global Environmental Monitoring. For many atmospheric scientists, Bob White was a welcome late addition to the U.S. delegation, and his emphasis on climate matched their own interest in the subject. And as a coordinating unit, Earthwatch initially managed to

“‘integrate’ information gathered from across the U.N. system” relatively effectively.

But it quickly became clear that Earthwatch was more of a bureaucratic clearinghouse for existing research efforts than it was a program to develop the kind of sophisticated infrastructure for scientific research called for by the atmospheric science community. Many scientists ultimately found the program disappointing. As an early

344 “Scope Paper.” 345 Walter Sullivan, “U.N. Parley Endorses Air Monitoring Net,” The New York Times, June 8, 1972, pg. 1.

153 foray into the problems of global environmental sustainability, Tom Malone later remembered, Earthwatch was “a worthwhile initiative.”346 But as the brainchild of diplomats and politicians rather than scientists, it also reflected certain “lack of imagination” in its approach to environmental research.347 SCEP, SMIC and SCOPE proposed broadening and deepening global scientific research through a creative new systems science methodology; Earthwatch merely sought to integrate existing research—albeit worthwhile research—already being conducted “within the U.N. system.”348 The U.S. delegation’s initiative incorporated a number of the specific recommendations for environmental monitoring contained in the SCEP and SMIC reports, but these improvements did not reflect any larger systematic changes. Train and his colleagues were careful to divorce their proposal from systems scientists’ larger global vision as best they could. It was, at best, a cautious proposal.

The Administration’s cautious, bureaucratic approach to Earthwatch extended to the delegations’ major financial initiatives as well. The U.S. delegation’s most important proposal at the Conference was a voluntary five year plan to raise $100 million dollars for the United Nations’ environmental programs, and it was called the U.N. Environment

Fund. The Administration designed the proposed Fund primarily to support research, education, and environmental monitoring within the international community, but the money was also meant to strengthen individual nations’ environmental management capabilities through regional-level environmental training and coordination.349 The

Administration specifically cautioned against allowing the fund to be diverted into

346 Interview with Tom Malone, June 4, 2009. 347 Ibid. 348 “Earthwatch,” UNEP’s website, http://earthwatch.unep.net/about/index.php. 349 Ibid.

154 specific intra-national economic or even environmental development programs, however.350 The Fund would provide the financial and bureaucratic support necessary for implementing Earthwatch, along with the U.S. delegation’s other specific institutional initiatives, but Train and his colleagues hoped to keep it separate from the pervasive geopolitical concerns over development and international aid.

In general, Nixon and his aides sought to streamline and publicize the U.N.’s existing American-sponsored environmental programs, but the President had no interest in empowering the U.N. by creating new or autonomous international agencies. Nowhere was this more prevalent than in discussions over administration of the U.N. Environment

Fund. The Fund both required and supported some sort of bureaucratic structure within the United Nations, and the U.S. delegation’s proposal for a “small coordinating unit” mirrored efforts to coordinate and streamline the environmental bureaucracy at home.

Within the United Nations, a variety of agencies, including the Food and Agriculture

Organization, the World Meteorological Organization, and the World Health

Organization, all worked on aspects of environmental monitoring, protection, and management.351 A number of Conference delegates felt that the United Nations should create an entirely new agency to deal with environmental problems.352 The Nixon

Administration disagreed, fearing that a separate agency would be costly and redundant.353 Instead, the U.S. favored a permanent environmental secretariat that would coordinate existing agencies’ environmental activities, much as the domestic CEQ coordinated the environmental activities of other federal and state agencies at home. The

350 “Scope Paper.” 351 Gladwin Hill, “Plan for Coordinating Unit Backed at Ecology Parley,” The New York Times, June 14, 1972; “Scope Paper.” 352 “Scope Paper.” 353 Ibid.

155 U.S. held the lion’s share of the Environmental Fund’s purse strings, and the Conference eventually backed the plan and created the United Nations Environment Program, an environmental secretariat based in Nairobi, Kenya, with Maurice Strong as its head.354

The U.S. delegation also supported three initiatives on specific international environmental issues that resonated with environmentalists at home. They included support for a convention on ocean dumping, a 10-year moratorium on whaling, and the establishment of a set of “World Heritage Sites,” modeled after the American National

Parks System.355 Each of these international initiatives spoke to widely held domestic concerns, and alongside Earthwatch and the Environment Fund, discussions about these issues at Stockholm gave the U.S. delegation an opportunity to capitalize politically on its environmental leadership both at home and abroad.356

354 Hill, “Plan for Coordinating Unit.” 355 The concern over whaling led to perhaps the most bizarre episode of the Conference, which New York Times science writer Walter Sullivan captured in a story called “The Cry of the Vanishing Whale Heeded at Stockholm.” In a field on the outskirts of Stockholm, the United States’ most prominent hippie commune, the Hog Farm Commune, hosted what was essentially a “whale-in” to protest whale hunting and offer support for the initiative on the 10-year whaling moratorium. The event, which entailed what Sullivan described as “the lugubrious cry of whales across the pine-studded landscape” generated by both tape recordings of actual whales and human imitations of whales, drew a surprising assortment of Conference delegates and government officials. In a particularly strange series of events, Maurice Strong gave a speech supporting the whaling moratorium, which was followed by a brief poem, “ostensibly written by a whale, with prolonged moans and groans.” Stuart Brand, the author of the award-winning 1971 The Last Whole Earth Catalogue, then mounted the stage in a “plumed top hat,” to introduce former Secretary of the Interior and occasional bugbear of environmentalists Walter Hickel. The next day, the Hog Farm hippies dressed one of their buses up in plastic bags to look roughly like a whale, and then followed it through the streets making whale noises and shouting the Swedish word for whale, “val!” Hickel, Strong, and the Hog Farm hippies all expressed their support for the moratorium, but in reality, the Japanese and the absent Soviets were the only groups that counted when it came to the whaling accord, because only the Japanese and the Soviets still engaged in large-scale commercial whaling in 1972. Argentina, Australia, Canada, Britain, France, Denmark, Iceland, Norway, Mexico, Panama, and South Africa did still take whales, and, along with Japan and the Soviet Union, constituted an international whaling commission. But the regulations mostly affected Japanese and Soviet whalers, and they ignored the United Nations’ agreement—and the Hog Farm hippies. The accord collapsed less than a month after the Conference. Walter Sullivan, “The Cry of the Vanishing Whale Heeded in Stockholm,” The New York Times, June 9, 1972; Walter Sullivan, “Whaling Halt Urged in Stockholm,” The New York Times, June 10, 1972; “Postscript to Stockholm,” The New York Times, July 7, 1972. 356 The convention on ocean dumping in particular was, in the words of Russell Train, “keyed to items on [Nixon’s] domestic legislative program.” In 1971, Congress challenged Nixon on his approach to water pollution with a unanimously supported bill to clean up America’s waterways by 1985. In some respects,

156 In Stockholm, however, the U.S. delegation’s achievements were quickly overshadowed by a second set of discussions about the global environment that the Nixon

Administration had explicitly hoped to avoid. Initiated by Sweden and China, and then taken up by developing nations, these were philosophical and diplomatic discussions that put environmental protection in the context of development, nuclear disarmament, and international peace. Much as Nixon’s aides attempted to address environmental issues in terms favorable to broader U.S. domestic and foreign policies, so too did U.N. members seek to redefine “the environment” to fit their own specific economic and political interests.357 And just as Nixon had feared, these interests frequently collided with the geopolitical priorities of the United States.

For the most part, the geopolitical concerns associated with the global environment were already on the table when the delegates arrived in Stockholm. Like many events held under the auspices of the United Nations, most of the major debates surrounding the U.N. Conference on the Human Environment occurred in the months and years leading up to the Conference itself. Four major preparatory meetings in 1971 and

’72 brought most of the key issues to the fore long before June of 1972. Drafts of the

“Declaration on the Human Environment,” set to be released at the close of the

Nixon’s ocean dumping legislation was an attempt to take the initiative on water pollution and the environment more generally back from Congress. In the President’s February “environmental message” to congress, Nixon had introduced similar domestic legislation, and Train explicitly modeled the Administration’s proposals for the international convention after this domestic legislation, hoping that the international convention would help support his proposals at home. John W. Finney, “Bill to Clean Up Waterways by ’85,” The New York Times, Nov. 3, 1971; Letter from Christopher DeMuth to John Whitaker, Nov. 3, 1971, John C. Whitaker Papers; “Talking Points for John Ehrlichman,” June 13, 1972, John C. Whitaker Papers. For more on Train and World Heritage Sites, see Nash, Wilderness and the American Mind, McCormick, Reclaiming Paradise, and J. Brooks Flippin, Nixon and the Environment (Albuquerque: University of New Mexico Press, 2000), and Conservative Conservationists: Russell Train and the Emergence of American Environmentalism (Baton Rouge: Louisiana State University, 2006). 357 Memorandum for the President from John C. Whitaker via John D. Ehrlichman, June 23, 1972, John C. Whitaker Papers.

157 Conference, were largely complete. The Founex Report had made it clear that the relationship between environmental protection and Third World development would be a central concern of many Conference participants. Negotiations over language dealing with nuclear weapons and weapons testing in the spring of 1972 presaged a discussion over disarmament in Stockholm, and tensions over the relationships between

“superpowers,” non-nuclear nations like Japan and Sweden, and the less developed countries of South America, Africa, and Asia promised to pervade discussions about obligations for international environmental protection.

For the U.S. delegation, however, there were a few important and somewhat unexpected wrinkles in Stockholm that undermined the Administration’s public relations goals for the conference. First, despite pointed and repeated attacks on the

Administration’s approach to the Conference in the months leading up to Stockholm by members of Congress like Claiborne Pell and Clifford Case and by representatives from

America’s environmental organizations, the Nixon Administration somehow seems to have been unprepared for the extent to which nations critical of American foreign policy used the environment as a way to attack the U.S. politically in the official Conference forum.358 Just as it was at home, the main issue in Stockholm was Vietnam. During his opening remarks on the first day of the Conference, June 6, Swedish Prime Minister Olof

Palme charged the U.S. with engaging in “ecological warfare” in Southeast Asia. “The immense destruction brought about by indiscriminate bombing, by large-scale use of bulldozers and herbicides is an outrage sometimes described as ecocide,” he declared,

358 It would be astonishing if Nixon’s aides truly did not see these attacks coming; it is more likely that they simply had no recourse to stop or even counter them at an international conference like the one at Stockholm. Still, there is very little discussion in the Conference files of Ehrlichman, Whitaker, Train, or Herter about how to best respond to these attacks, and not until after the opening statements do they begin to discuss how to handle the public relations problems these attacks engendered.

158 echoing similar statements made by Claiborne Pell on the Senate floor in the previous months.359 Palme framed the United States’ destruction of the Vietnamese environment in terms of a larger relationship between armaments and environmental destruction. He called on the Conference to urge an end to large-scale arms production worldwide. The

Chinese delegation put an even finer point on the Swedes’ criticism of the Vietnam War.

In his first major pronouncement on June 10, Tang Ke, the head of the delegation, asked the Conference to “strongly condemn the United States for their wonton bombings and shellings, use of chemical weapons, massacre of the people, destruction of human lives, annihilation of plants and animals, and destruction of the environment.”360

Similar criticisms came from individuals outside of the official Conference in

Stockholm participating in an informal “Environment Forum,” where prominent environmental activists and NGOs voiced the myriad environmental concerns of the various constituents not represented at the U.N.361 The Forum included almost 500 accredited non-governmental organizations (NGOs)—many of them American—and scores of prominent individuals, including some official Conference observers and

359 Gladwin Hill, “Draft Calls for Ecological Responsibility,” The New York Times, June 7, 1972. 360 Gladwin Hill, “China Denounces U.S. on Pollution,” The New York Times, June 10, 1972. 361 In addition to his penchant for calling the U.N. Conference a “parley,” Gladwin Hill of The New York Times in particular referred to the Environment Forum as a “counter-conference.” The name is in some senses misleading as the Forum was meant to complement the U.N. Conference, but in others, it actually captures the spirit of the Forum quite well. Especially among American youths, many environmental activists questioned the sincerity and effectiveness of a conference designed by mid-level diplomats and bureaucrats, for mid-level diplomats and bureaucrats. The Forum gave skeptics a voice, if only an unofficial one. In addition, because the Forum drew from a non-governmental pool, it included many elements of the popular “counterculture” often associated with environmental awareness and protection in the U.S. The Hog Farm Commune hippies—youths who lived mostly in buses and had in the past served as a security force under the leadership of their founder, Wavy Gravey—provide an apt example of just how close Rocky Mountain Center for the Environment official Roger Hansen’s description of the Forum as “a sort of environmental Woodstock for the world’s affluent youth” actually came. In that sense, it was in fact a “countercultural” conference. Hill, “Sense of Accomplishment Buoys Delegates Leaving Ecology Talks,” The New York Times, June 18, 1972; Walter Sullivan, “The Cry of the Vanishing Whale”: “Statement of Roger P. Hansen before the Secretary of State’s Advisory Committee on the United Nations Conference on the Human Environment,” March 7, 1972, John C. Whitaker Papers.

159 delegates.362 Participants espoused a wide range of objectives and diverse viewpoints, but most opposed the war in Vietnam and supported binding limitations on nuclear weapons in one way or another. They hosted a major protest of the Vietnam War in Stockholm during the second week of the Conference.363

Perhaps more surprising than the invective hurled at the U.S. for its involvement in Vietnam was China’s unexpectedly aggressive overall anti-Western position at the

Conference. Sino-American relations had thawed remarkably under Nixon, and the

President had visited the communist nation in February of 1972. But Ke apparently hoped to use the Conference to establish the People’s Republic of China as a major force in the U.N., and he and his colleagues staked out a position as the champions of the Third

World in the face of the “imperialistic superpowers.”364 The PRC had not yet been admitted to the U.N. during the first of the Conference preparatory meetings in 1971, and

Ke insisted upon reopening the all but finished “Declaration on the Human Environment” for “more democratic” revision now that the PRC had been seated.365 Once reopened, negotiations over the language of the Declaration became a forum for criticism and dissent, with the U.S. as a primary target. Despite the concessions of the Founex

Document, many less developed countries still harbored deep reservations about a U.N. environmental effort that they felt didn’t sufficiently account for their development needs. China fueled their discontent. “We are firmly opposed to the superpowers

362 The number comes from a memo for the President from Russell E. Train, June 19, 1972, John C. Whitaker Papers. 363 “Talking Points for John Ehrlichman,” June 13, 1972. 364 In their controversial Mao: The Unknown Story (New York: Random House, 2005), Jan Chung and Jon Halliday argue that the Chinese sought to court the Third World in order to undercut the power of the Soviet Union during this period. Their behavior at the Conference would seem to bear this out. Chung and Halliday also argue that one of Mao’s ultimate objectives—in fact, an obsession—was for China to become a nuclear power on par with the U.S. and Soviet Union. China’s resistance to disarmament language at the Conference reflected an active nuclear testing program at home. 365 Hill, “China Denounces U.S.”; “Sense of Accomplishment.”

160 subjecting other countries to their control and plunder on the pretext of improving the human environment,” Ke declared. 366 Some African nations, Algeria most vocal among them, had begun to demand compensation for the historical exploitation of their environments by colonial powers, and Ke supported these demands.367 “Victim countries,” as Ke called them, “have the right to apply sanctions against and demand compensation from the culprit countries.”368 Because of its focus on the interests of less developed countries, the Chinese delegation’s critical remarks drew heavy applause from an audience whose Third World representatives outnumbered the developed world by more than two to one.369

The U.S. delegation handled these developments rather badly. Tang Ke’s scathing opening remarks proved particularly damaging, not least because Russell Train and his colleagues had to clear any unscripted statements with the State Department, the

Department of Defense, and the White House before offering a rebuttal. The series of events was embarrassing. The U.S. delegation asked for rebuttal time immediately after

Ke finished speaking, but after nearly five hours of deliberation, Christian Herter of the

State Department had to ask that the rebuttal be deferred for the weekend. As an alternative to an official rebuttal, he called a news conference for that afternoon, but he had to cancel at the last minute because the delegation hadn’t come up with an acceptable message for the press.370 Train’s rebuttal, finally delivered the following Monday, only led the press to revisit China’s criticisms, and it contained little of substance. As it happened, the Chinese had already walked out of the Conference during a speech by

366 Hill, “Sense of Accomplishment.” 367 Hill, “Draft Calls for Ecological Responsibility.” 368 Hill, “China Denounces U.S.” 369 Ibid. 370 Ibid.

161 delegates from what they called the “puppet clique” of South Vietnam before Train had a chance to present his case.371 A photo of child star turned diplomat Shirley Temple Black applauding Train showed row upon row of empty seats.372

Conclusion

It is easy to lament, as Russell Train did in his 1972 evaluation of the Conference on the Human Environment, the “politicization” of the global environment that occurred at Stockholm. In the years since, many scientists studying climate and the global environment have echoed Train’s sentiment, complaining that the chicanery of politics has continually invaded upon rational, science-based plans for environmental management and protection. But it is a misguided lament. The global environment as a concept has always been political, thanks in large part to the scientists who first helped to describe it. SCEP and SMIC, after all, were undertaken with the expressed intent of informing a nascent political discussion about the global environmental crisis; The Limits to Growth was nothing if not politically prognosticative. Scientists’ idea of a world made

371 Hill, “Plan for Coordinating Unit.” The “puppet clique” comment was reported in Hill, “China Denounces U.S.” 372 The photos were from the Associated Press International, and appeared alongside Hill, “Plan for Coordinating Unit.” The Administration tried to counter the negative press by launching a public relations campaign during the second week of the Conference. Even the campaign’s designer, John Whitaker, expected it to fall flat, and it did. Whitaker called John Ehrlichman on June 13 to brief him on the first half of the Conference. Despite “some solid accomplishments led by the United States,” Whitaker lamented, “the news leads for the next few days will inevitably be negative.” He was not sanguine about press coverage of Train’s rebuttal to the Chinese, and feared that a Vietnam protest associated with the Environment Forum would dominate the following day’s headlines. Indian Prime Minister Indira Gandhi, with whom Nixon had a thorny relationship, was scheduled to speak Wednesday. She would “not be any help” to the American cause. Whitaker tried to counter these leads with diversions. He scheduled daily briefings with Conference delegates for the U.S. press, and tried to play up public relations stunts staged by Train and EPA chief William Ruckelshaus. Ruckelshaus visited a recycling plant. Train, alongside members of the Chinese delegation, planted Chinese Elm trees grown in the U.S. in a park in Stockholm. The efforts were ineffective (in fact, they drew ridicule), but as Ehrlichman noted, they were “about the best he [Whitaker] could do.” “Talking Points for John Ehrlichman,” June 13, 1972, John C. Whitaker Papers.

162 up of complex, interrelated large-scale systems helped to buttress Maurice Strong’s “new globalism” in international politics by highlighting mutual interest and displaying the necessity for international cooperation in dealing with global problems. Scientists themselves explicitly framed their approach in these internationalist terms. It is unfair to impugn the actual scientific work involved in these reports as somehow biased or non- objective, but the projects themselves arose self-consciously out of a desire to put the global environment on the political agenda in the 1970s, and it would be naïve to ignore the political motivations of their participants.

For the scientific community, perhaps a more honest lament than Train’s complaint about the politicization of the global environment would be that the scientists who studied and described its large-scale systems did not play a greater role in defining its politics. Alongside the U.N. leaders they supported, system scientists found that their vision for a form of cooperative scientific and political globalism built around common world-scale environmental problems stood in tension with an existing geopolitical system that privileged national and regional scale economic concerns. The regime of global environmental governance established at Stockholm ultimately reflected the interests of these established constituencies at least as much as it did a collective concern for humanity’s impact on the whole Earth.

Then again, maybe this wasn’t necessarily such a bad thing, especially in the short run. After all, a new framework for global environmental politics was established at

Stockholm, and it was established with some success.373 As Anthony Lewis of The New

373 This success was in some senses quite remarkable. In Seeing Like a State: How Certain Schemes to Improve the Human Condition Have Failed (New Haven: Yale University Press, 1998), James C. Scott demonstrates the ways in which governments, often backed by scientists and other management experts, have attempted to impose rational order upon nature and space as a way to gain control over populations

163 York Times reported in his recap of the event, “One Confused Earth,” the very occurrence of the Conference—especially after the repeated confrontations between East and West, rich and poor, Communist and Capitalist in the preceding months—represented a victory for international environmental consciousness. The event drew more than a hundred nations together to discuss a relatively new set of disconcerting global issues for the first time, and it achieved a good deal of real political success.374 Member states established a number of benchmark regulations for specific issues including ocean dumping, whaling, and toxic wastes, and it provided an ambitious set of 27 guiding principles on which to base future international environmental agreements. The meeting also established the machinery for developing and implementing new international environmental policy within the new United Nations Environment Program, based in Nairobi, Kenya.375

and their resources. As Vandana Shiva argues in her essay “The Greening of the Global Reach,” the reordering of local and regional environmental problems into a “global environment” similarly provided First World scientists and governments a way to strengthen the international community’s control over the landscapes, water and fossil fuel resources, and populations of individual states, particularly those of the less developed world. As Scott points out, however, large-scale social engineering has historically involved more than simply a reordering of nature or society; it has also required a commitment to a “high- modernist” ideology, an authoritarian state, and a weak civil society willing to accept direction from a centralized authority. As it set out to define the “global environment,” the United Nations found itself struggling with each of these requirements. Environmentalism itself was in part a challenge to “high- modernist” ideology, and to the agricultural and industrial development it demanded. Moreover, as a loosely-knit and non-binding federation of sovereign states, the United Nations sought, as one of its primary missions, to undercut the dictatorial actions of individual authoritarian states. Even if the organization could agree on a global vision of environmental stewardship, it could hardly force its members to submit to that vision. And finally, with an angry flowering of revolutions and independence movements in full swing throughout the Third World—and with an increasingly complex Cold War dominating international politics—the civil societies of the world were anything but submissive to quasi-colonial Western control in the 1970s. As a result, the ordering of the new global environment was a contested and deeply political process of back-and-forth negotiations that included continued challenges to the authority and priorities of the Western politicians, bureaucrats, and scientific experts at the heart of international environmentalism. That the U.N. established a regime of environmental management at all under these circumstances is, as Anthony Lewis of The New York Times pointed out at the time, quite impressive. Anthony Lewis, “One Confused Earth,” The New York Times, June 17, 1972; Vandana Shiva, “The Greening of the Global Reach,” Global Ecology: A New Arena for Political Conflict, edited by Wolfgang Sachs (London: Zed Books, 1993): 149-56. 374 Lewis, “One Confused Earth.” 375 Hill, “Plan For Coordinating Unit.”

164 The Conference made few inroads into the dominance of the interests of nation- states in international politics, but the concept of non-renewable “world resources” began to resonate with delegates from both developing and developed nations. Growing Third

World political interest helped to buoy First World environmental initiatives, both within

UNEP and within related U.N. agencies like the Food and Agriculture Organization and the World Health Organization. By incorporating developing nations and their concerns over the social and political ramifications of environmental degradation and protection into an international political process, the U.N. Conference on the Human Environment helped to establish the global environment as an increasingly important issue of geopolitical interest.376

The scientific community may have been overshadowed at the Conference by the development issue, but they didn’t exactly lose out at Stockholm either. Despite concerns over nations’ sovereignty and security, “Earthwatch” reinforced scientists’ efforts to monitor and study the environment at a global scale, especially within the

World Meteorological Organization. UNEP promised to help focus the international environmental policy-making process in a way that would allow policy-makers better access to scientific information, and media coverage of the Conference helped stir up greater public interest in international environmental issues involving the oceans and the atmosphere. It was hardly the New Atlantis that some had quietly dreamed of during the

SCEP and SMIC meetings, but it was a start.

376 In Hays’s words, this “stress on efficient development and use of material resources such as water, forests, and soils known as the conservation movement” described a type of environmental protection that played out in the United States in the first four decades of the twentieth century—decades that he chronicles in his other foundational work on American environmentalism, Conservation and the Gospel of Efficiency: The Progressive Conservation Movement, 1890-1920 (Cambridge: Harvard University Press, 1959). The quotation is from Hays, Beauty, Health, and Permanence, 4.

165 As the 1970s wore on, however, atmospheric scientists found that the political framework for global environmental governance established at Stockholm, dominated as it was by national and regional interests, made it difficult to gain support within the

United Nations for the truly global environmental problems involving the

atmosphere—ozone depletion, aerosols, and CO2-induced climate change.

At first, climate took a back seat to other issues because it was as yet a novel problem that could barely even be described with any kind of certainty, let alone incorporated into an international system of environmental protection. In 1972, there were still significant divisions within the scientific community over the true causes, extent, and even direction of climatic change.377 Bob White agreed that the WMO should study the problem, but despite the urgings of SCEP, SMIC, and Global Environmental

Monitoring, it was not a primary focus of the international environmental community.

Soon, though, as the potential impacts of rising levels of atmospheric CO2 became clearer, it also became increasingly clear that the United Nations Conference on the

Human Environment, for all its globalizing rhetoric, had no way of dealing with a truly global environmental problem like climate change. In Stockholm, scientists and political leaders framed everything from waste management to soil depletion to land use change to water pollution as parts of a larger, global environmental crisis. But while these problems may have crossed borders and represented a cumulative or aggregate threat to the global environment, they nevertheless existed primarily within and between pre- existing political spaces, and these spaces involved particular people with particular

377 For more on the cooling/warming debate, see chapter 4.

166 interests. Pitched as global, they all also fit, albeit sometimes awkwardly, into an international politics centered upon the economically interested nation-state.

The global atmosphere, by contrast—fluid, dynamic, and borderless—had no such attachment to local or regional politics. Changes in climate or atmospheric composition might impact environments anywhere and/or everywhere, and in 1972 the interests at stake were still unclear. In a battle to define the global environmental in terms of competing development and natural resource interests, a global-level phenomenon not rooted in any specific national or even regional space had no interested constituency. So while climate change certainly fit into scientists’ vision for a cooperative global research effort administered by a revitalize United Nations, it remained a relative non-issue in international environmental politics until the 1980s. Only after another decade of effort would scientists succeed in making climatic and atmospheric change relevant to the regional and national scale interests at the heart of international politics.

167 Chapter 4 Climate, “The Environment,” and Scientific Activism in the 1970s

The SST controversy and the U.N. Conference on the Human Environment presaged the growing involvement of atmospheric scientists in global environmental issues during the 1970s. After Congress killed the SST’s funding in 1971, many of these atmospheric scientists joined scientists from other disciplines also interested in the relationships between climate and natural resources to form a new scientific field: climate science. Throughout the 1970s, this loose, mutli-disciplinary scientific community worked to expand climatic research and to establish a place for climatic variability and climate change on national and international policy agendas. By 1980, climate scientists had collectively identified most of the major scientific components of the current, 21st century issue of global warming.

Climate scientists gained political influence in the 1970s despite significant divisions within their ranks. Individuals within the community disagreed over some of the basic conclusions of their research, as well as over the proper relationship between science and politics. Their disputes often reflected differences in the methodologies used by different disciplines and at different institutions. In the 1970s, the historical-statistical evidence of traditional climatologists pointed to a potential —perhaps an imminent ice age—caused by variations in solar radiation and by reflective airborne particles. At the same time, many atmospheric models began to show more distinctly that

increased CO2 would lead to global warming. These scientific differences became entwined with debates about how climate scientists presented their conclusions, which in turn involved disagreements over scientists’ role in making policy. Scientists who saw the threats from either cooling or warming as particularly imminent sought to incorporate

168 their work into public policy immediately. They often called for drastic measures to curb energy use and build more flexibility into vulnerable resource infrastructures. More conservative scientists objected to politicizing a scientific discussion still rife with uncertainties. They usually promoted further research in order to establish a scientific consensus. Theirs was a “wait and see” approach.

Climate science was born out of a concern for the environment, but climate scientists generally defined “the environment” differently than American environmentalists did. American environmentalists continued to focus largely on

Americans’ day-to-day quality of life. Proper waste management, clean air, ample local parks and open spaces, and access to the nation’s pristine wilderness areas increasingly defined the “good life” in America, and environmentalists saw degradation from industrial pollution and irresponsible development as a threat to these “back yard” environmental amenities.378 Climate scientists, on the other hand, addressed the environmental crisis primarily in terms of the food, water, and other natural resources essential to life itself. They sympathized with the broad, humanitarian goals of the international environmental movement, articulated at the 1972 Stockholm Conference.379

Climate scientists hoped to use their expertise to help mitigate the detrimental effects of climatic variability on development and agriculture, and to better plan for the potential impacts of long-term climatic change on human environments, domestically and abroad.

378 Samuel P. Hays’ Beauty, Health, and Permanence: Environmental Politics in the United States, 1955- 1985 (New York: Cambridge University Press, 1987) still provides the most comprehensive analysis of post-war American environmentalism. See also Robert Gottlieb, Forcing the Spring: The Transformation of the American Environmental Movement (Washington, D.C.: Island, Press, 1993); Hal K. Rothman, The Greening of a Nation?: Environmentalism in the United States Since 1945 (Fort Worth, TX: Harcourt Brace & Company, 1998); Kirkpatrick Sale, The Green Revolution: The American Environmental Movement, 1962-1992 (New York: Hill & Wang, 1993); Philip Shabecoff, A Fierce Green Fire: The American Environmental Movement (revised edition) (Washington, D.C.: Island Press, 2003). 379 The U.N. Conference on the Human Environment is discussed extensively in Chapter 3.

169 More akin to the expert resource managers of the Progressive Era conservation movement than the defensive environmentalists of their own day, climate scientists worked within government to help policy-makers create rational policies for energy, development, and the environment on a global scale.

Environmentalists initially responded to the issue of climate change tepidly. They hesitated to rally behind a loosely defined community of scientists whose definitions of

“the environment” differed from their own, and whose specific political goals remained vague. Climate scientists chose to work within the very agencies environmentalists sought to challenge, and they usually did so with few recognizably “environmental” objectives.

Perhaps most importantly, the nature of climate change itself made it difficult for professional environmentalists to incorporate the issue into their national and international political strategies. Members of groups like the Sierra Club and Wilderness

Society took an interest in international environmental issues, but they selected their campaigns carefully. They looked for clear problems with definitive solutions. Climate change was highly technical, global in scale, rife with scientific uncertainties, and it occurred over the course of decades and centuries. No obvious political or legal structures existed to implement solutions to the problems of climate, and few, if any, solutions presented themselves. Until the mid-1980s, climate change remained almost entirely the purview of the scientific community.

170 Climate Science and Science Funding in the 1970s: A View from NCAR

The terms “climate science” and “climate scientist” didn’t appear in the pages of journals like Science until as late as 1976, but a community of atmospheric scientists and climatologists had begun to engage in a broad, multi-disciplinary discussion about climate change and its implications as early as 1970.380 As with many scientific endeavors, the young field was shaped in part by its funding. The early 1970s saw the first overall decrease in funding for natural science in more than a decade, and the

National Science Foundation (NSF), the Departments of Commerce, Defense, and

Transportation, and the National Aeronautics and Space Administration—the primary sources of funding for atmospheric and climatological research—began to favor specific projects that applied science “in the national interest.”381 In response, scientific institutions and individual university scientists began to look toward climate science as a new application of their basic research, relevant to politically salient concerns over the health of the environment. Climate science thus came to maturity as an “environmental” science intrinsically tied to concerns over the environmental impacts of climate change.

380 Science, “Speaking of Science: Concern Over Climate: Researchers Increasingly Go Public,” (Vol 192, Issue 4236), April 16, 1976, pg. 246-247, cited in “climate science, n.” OED Online. June, 2010. Oxford University Press, 2010. http://dictionary.oed.com/cgi/entry/50041503/50041503se13. 381 In unadjusted dollars, the total federal science budget never actually decreased, but between 1972 and 1973, inflation far outstripped the growth in federal science funding. Overall funding for research hovered around $5.5 billion over these those two years, increasing by less than 1%; average annual inflation for the 1970s was around 6%. The total expenditure for atmospheric sciences, meanwhile, did decrease in unadjusted dollars, dropping from around $259 million in 1971 to $241 million in 1972, and then again from $302 million to $289 million between 1975 and 1976. The National Science Foundation, chief source of Federal funding for atmospheric science, had its overall budget trimmed in terms of growth, but like the Federal Government more generally, the NSF never suffered an actual “decrease” in unadjusted dollars. NOAA and the Department of Commerce, by contrast, incurred actual cuts, made all the more severe by the decade’s inflation. National Science Foundation, Federal Funds for Research and Development Detailed Historical Tables: Fiscal Years 1956-1994, NSF 94-331 (Bethesda, MD: Quantum Research Corp., 1994). See also Joseph Paul Martino, Science Funding: Politics and Porkbarrel (New Brunswick, NJ: Transaction Publishers, 1992); James Savage, Funding Science in America: Congress, Universities, and the Politics of the Academic Pork Barrel (New York: Cambridge University Press, 1999).

171 The effects of the drive for socially relevant research on climate science were particularly pronounced at the National Center for Atmospheric Research (NCAR) in

Boulder, Colorado. Founded in 1960 as the institutional hub of a consortium of universities conducting atmospheric research, the University Corporation for

Atmospheric Research (UCAR), NCAR had by the late 1960s become a premier institution of atmospheric science. At the outset of the decade, the NSF estimated that

NCAR represented about half of its overall expenditures on research into the atmosphere.382 In the 1970s, the institution began to apply more and more of these resources to the study of climate.

Hired as a post-doc in 1971, Stephen Schneider played a key role in establishing climate change as an official programmatic focus at NCAR. Alongside SCEP and SMIC authors William Kellogg and Philip Thompson, Schneider took advantage of NCAR’s

NSF-mandated institutional restructuring in order to promote an integrated, multi- disciplinary approach to climate science.

Schneider was drawn to climate science as much for its social and political relevance as for its intellectual challenges. Politically, Schneider was a liberal, but he had reservations about both the Democratic Party 'establishment' and the new left radicals who challenged it. As a graduate student at Columbia University in 1968, Schneider had immersed himself in the campus’s tumultuous political scene. He won a seat as the engineering department’s representative on the student council, where he tried to walk the line between the “inappropriate methods” of the radical Students for a Democratic

382 National Science Foundation Advisory Panel for Atmospheric Research, “Report on the National Center for Atmospheric Research,” November, 1970, pg. 1, UCAR/NCAR Archives.

172 Society and the business-as-usual program of the student council.383 A disappointed

McCarthy Democrat, he opposed the Vietnam War and saw a real need for political change in America, but he hoped that change could be driven by reason, moderation, and science rather than brick-throwing and revolution. 384 Schneider came to NCAR in part because he believed in NCAR founder and then UCAR director Walter Orr Roberts’s commitment to using science to benefit society.

He also found climate modeling fascinating. In 1970, while still at Columbia,

Schneider began to work with a group of modelers at NASA’s Goddard Institute for

Space Studies (GISS). Originally trained in plasma physics, he worked on climate models as a way to prolong his Ph.D. program past his 26th birthday and thereby avoid the draft. “It was absolutely exciting to me,” Schneider recalled three decades later,

“that I could sit down at a key punch, type up a box of cards and hold in my hands

the capacity to simulate the earth…how exciting it was that you could actually

simulate something as crazy as the earth, and then pollute the model, and figure

out what might happen…”385

Schneider was good at modeling. In early 1971, he collaborated with GISS’s

Ishtiaque Rasool to demonstrate the overall cooling effect of pumping industrial aerosols into the Earth’s atmosphere. Their controversial study found that a four-fold increase in

background aerosol scattering could overwhelm CO2-induced warming, and, if sustained,

383 Interview with Stephen Schneider by Bob Chervin, January 10-13, 2002, AMS/UCAR Tape Recorded Interview Project. 384 Schneider recalls canvassing for McCarthy in a labor district of Connecticut with two Asian-American colleagues on the day Johnson announced his intention not withdraw from the Presidential race (March 31, 1968). Schneider had “nice conversations with most families” he visited, but his colleagues received “chilly receptions from white laborites.” Stephen H. Schneider, e-mail correspondence, 10/9/08. 385 Interview with Stephen Schneider by Bob Chervin, Jan10-13, 2002.

173 “be sufficient to trigger an ice age.”386 Although Schneider and Rasool soon abandoned the idea that negative forcing of aerosols could overpower or counteract the positive forcing of greenhouse gases, in 1971 the paper made a splash in the climate change community. It won Schneider—whom the senior Rasool sent out to present the paper at a series of conferences shortly after its publication in Science—both respect and notoriety.387

As brash and outspoken as he was bright and talented, Schneider introduced himself to NCAR’s William Kellogg during a conference shortly after the release of the

SCEP report. Kellogg invited the eager 26-year-old to work as the “rapporteur” for

SMIC during the summer of 1971.388 When Kellogg suggested that Schneider continue his work on aerosols and climate in a one-year post-doc at NCAR’s Advanced Studies

Program—a sort of clearing house for talented post-docs generally interested in atmospheric science—the young scientist jumped at the opportunity.389

Schneider made it his business at NCAR to use his skills of political organization to expand the scope of the institution’s discussions about climate change. He framed the

386 As Schneider often points out, the paper did not actually predict an ice age, but merely demonstrated the possibility that aerosols might create the conditions for the development of another ice age. Interview with Stephen Schneider by Bob Chervin, Jan 10-13, 2002; S.I. Rasool and S.H. Schneider, “Atmospheric Carbon Dioxide and Aerosols: Effects of Large Scale Increases on Global Climate,” Science, New Series, vol. 173, no. 3992 (July 9, 1971): 138. 387 Schneider recalls his behind-the-scenes role at the SMIC conference with some amusement. Told in no uncertain terms to defer to senior scientists and generally keep his mouth shut, Schneider kept his work with Rasool to himself. During the conference, however, the International Herald Tribune ran an article carrying an interview with Rasool about the work, and claiming that Schneider had come to Stockholm to present their research. With his name in the Tribune attached to a reference to the coming of an ice age, Schneider, predictably, became the center of attention for the press conference the next day, quickly earning him a reputation as a very pubic scientist that he would live up to throughout his career. Interview with Stephen Schneider, Jan 10-13, 2002; Chicago Tribune, “Ice Age Around the Corner,” July 10, 1971. 388 The term “Rapporteur,” derived from the French, is used mainly in international and European legal and political settings, and refers to a person appointed as an investigator or reporter for a deliberative body. For one reason or another, scientists chose to use the term to describe the lowly position of compiling, editing, and distributing reports and papers for the SMIC and SCEP conferences. The origin of the term as used by scientists—and of the position—merits further research by an interested historian of science. 389 Interview with Stephen Schneider by Bob Chervin, Jan 10-13, 2002.

174 conclusions of atmospheric physicists and global circulation modelers in terms of their ecological, economic, and social impacts, and advocated a stronger program of interdisciplinary climate science at the institution.390 With encouragement from Kellogg and Thompson, as well as from his mentors in climate modeling Warren Washington and

Bob Dickinson, Schneider set up an informal series of afternoon “wine and cheese” talks on controversial climate-related topics, called the “Climate Club.”391 Paul Ehrlich, a well-known Stanford population biologist worried about human population growth, and J.

Murray Mitchell, Jr., a controversial University of Washington climate expert working for NOAA who warned of rapid transitions between glacial and interglacial periods, drew large audiences.

In 1972, Schneider began to drum up support within the institution for a new climate research project.392 The NSF and UCAR had set up a “Joint Evaluate

Committee” (JEC) to redefine NCAR’s scientific priorities and shake up its management the previous year, and Schneider saw an opportunity to push climate science as an institutional focus. Schneider, Kellogg, Dickinson, and Thompson submitted an official proposal to for a major NCAR climate initiative in 1973. The JEC not only approved the proposal, it ranked climate change at the top of the agenda for new institutional projects.

Schneider—half way through a one-year post-doc—became the project’s deputy director

390 Schneider describes his efforts to popularize climate science at NCAR as an attempt to demonstrate the “market pull” of the research within the institution. Given Schneider’s political orientation in 1972, and given the emphasis on government regulation in his popular 1976 book The Genesis Strategy: Climate and Global Survival (New York: Delta, 1976), it seems likely that this characterization is quite an anachronism. Schneider had not yet become an advocate of market solutions (he would become one), and had little experience dealing with these types of principles—even metaphorically—in 1972. Interview with Stephen Schneider by Bob Chervin, Jan 10-13, 2002. 391 Interview with Stephen Schneider by Bob Chervin, Jan 10-13, 2002. 392 Ibid.

175 that winter.393 Schneider’s climate project ensured him—and NCAR—a place in an ongoing scientific debate about the nature of climate change.

The Disciplinary Landscape of Climate Science

In the early 1970s, the causes, extent, and even direction of climate change were hardly certain. Scientists from different disciplines approached the study of climate with different and sometimes competing methodologies, and they often disagreed over the severity and direction of a potential climatic change. In the early 1970s, the climate science community was characterized as much by these disciplinary and methodological divisions within a constant struggle for funding as it was a by a common set of concerns over climate change and the human environment.

Climate scientists generally agreed that climatic change could adversely affect the human environment, but their research produced two conflicting scenarios. The first, espoused primarily by atmospheric scientists working with advanced computer models, involved an overall increase in global temperatures caused by increases in atmospheric

CO2 and other greenhouse gases. The second, promoted largely by climatologists and geologists working both with models and with physical and documentary evidence of past climatic shifts, consisted of a relatively sudden decrease in global temperatures caused by various potential combinations of volcanic dust, industrial pollutants, land use changes, and the long-term cycles of solar radiation.394 Divisions between atmospheric

393 Ibid. 394 As Thomas C. Peterson of the National Climatic Data Center recently pointed out in the Bulletin of the American Meteorological Society, comparatively few climate scientists bought into the cooling scenario, and scientific papers predicting warming outnumbered those predicting cooling more than six to one during the 1970s. Peterson uses his analysis of journal articles from this time period as a way to debunk contemporary skeptics who use the global cooling theory to undercut the credibility of the overwhelming

176 scientists and historical climatologists were not hard and fast, however. Atmospheric models at different levels of sophistication also sometimes disagreed, alternately

predicting warming or cooling. CO2-induced warming remained the most likely scenario, but as late as the mid-1970s, scientists still hadn’t completely ruled out the cooling problem.

Atmospheric scientists, whose models highlighted the warming effects of

atmospheric CO2, dominated the study of climate in the late 1960s and early 1970s. As early as 1967, Sukiro Manabe and Richard Weatherald, both working at the Geophysical

Fluid Dynamics Laboratory in Princeton, had used numerical models to confirm the

longstanding theory that increases in atmospheric CO2 would lead to increases the global mean temperature.395 A decade of measurement at Mauna Loa, Hawaii, in Antarctica, and elsewhere firmly established the “Keeling Curve”—the upward sloping saw-toothed

line representing the amount of CO2 in the atmosphere based on David Keeling’s original

majority of scientists who now predict global warming. And, for the most part, his article hits the mark. But what Peterson fails to address in his article is the distinction between consensus as scientists envision it today and prevalence or popular credibility, which can exist independently of scientific consensus. It is true that comparatively few scientists espoused cooling as the most probable scenario, but many of those who did—like Reid Bryson—were particularly well-respected. Moreover, the media found it much easier to play up the threat of a new ice age than they did to imagine an unpredictable, warmer world, and as a result, the general public was perhaps equally if not more aware of the threat of cooling as they were of the threat of warming. As Peterson notes, there certainly was no consensus on global cooling in the 1970s, but that does not mean, as Sid Perkins has recently argued in ScienceNews, that cooling did not represent a credible scientific theory during at the time. That scientists seriously considered cooling is not a “myth” propounded by today’s global warming skeptics; many important scientists, including Bryson and Stephen Schneider, did in fact subscribe to the cooling hypothesis. Perkins’ primary complaint holds water, however. As he argues, the debate over cooling in the 1970s is an episode in the early history of climate science that global warming skeptics and deniers have willfully and invidiously misinterpreted to serve their own purposes. Sid Perkins, “Cooling Climate ‘Consensus’ of 1970s Never Was,” ScienceNews, vol. 174, no. 9 (October 25, 2008), 5-6. Thomas C. Peterson, William M. Connolley, and John Fleck, “The Myth of the 1970s Global Cooling Scientific Consensus,” in Bulletin of the American Meteorological Society, Volume 89, Issue 9 (September 2008), 1325-1337. 395 GFDL was then part of the U.S. Weather Bureau. In 1970, the Weather Bureau was folded into a new, broader institution overseen by the Department of Commerce, the National Oceanic and Atmospheric Administration (NOAA). GFDL became part of NOAA. Sukiro Manabe and Richard T. Weatherald, “Thermal Equilibrium of the Atmosphere with a Given Distribution of Relative Humidity,” Journal of the Atmospheric Sciences, vol. 24(3), 1967, 241-259.

177 measurements in 1957. CO2 had, as Keeling and Roger Revelle had predicted, increased significantly since 1957. In the Manabe-Weatherald model, if the increase continued to the point of doubling, scientists could expect to see an approximately 2.3°C temperature increase by the turn of the century, all other things being equal.396

But all other things were not equal. Meteorologists’ day-to-day temperature measurements revealed that although the early 20th century had experienced a modest overall warming, in the years between 1940 and 1970, the Earth had actually cooled. The

20th Century, climatologists recognized, had seen remarkably stable climatic conditions that stood in stark contrast to the variability of the geological past. Some feared that the

Earth’s string of interglacial good luck was finally coming to a close, possibly with help from its human inhabitants. At a conference of the American Association for the

Advancement of Science on the “Global Effects of Environmental Pollution” in Dallas in

1968, Reid Bryson of the University of Wisconsin explained that this seemingly anomalous cooling could be attributed to particles of smoke and dust in the atmosphere coming not just from volcanoes—which many climatologists agreed could be agents of climatic change—but also from human activities, most notably the slash/burn agricultural practices of less developed countries in the tropics.397

In the early 1970s, with temperature trends clearly showing a cooling in the

Northern Hemisphere from 1940 to 1970, Bryson and other historically-minded climatologists—a group that consisted of geologists, paleontologists, anthropologists,

396 This prediction—a 2.3°C rise in temperature—would, despite the relative simplicity of the Manabe- Weatherald model compared to later, higher-resolution climate models, remain within the a few points of the “best guess” of the ranges of modelers’ predictions for the next three decades (see Chapter 4 and 5). For more on Keeling and the Keeling Curve, see Chapter 1. See also Spencer Weart, The Discovery of Global Warming (Cambridge: Harvard University Press, 2003), 35-38. 397 Reid A. Bryson and Wayne M. Wendland, “Climatic Effects of Atmospheric Pollution,” Global Effects of Environmental Pollution, edited by S. Fred Singer (New York: Springer-Verlag, 1970), 130-138.

178 historians, and meteorologists—mustered convincing evidence from ice cores, tree rings, peat bogs, and historical documents that the Earth might, for one reason or another, be returning to the cooler (and leaner) times of previous centuries, if not to another full- fledged ice age. Basing their analyses and predictions on a combination of statistical climatic trends and human historical examples, this interdisciplinary group—Bryson the most outspoken among them—focused on the vulnerability of marginal agricultural areas to natural and anthropogenic climatic variation. Cyclical variances in solar radiation and increases in volcanic activity seemed to many to be the prime movers of this overall cooling, but Bryson and others also warned that human activities—including overgrazing, slash/burn agriculture, industrial pollution, and transportation emissions—threatened to speed up these natural processes by inadvertently modifying the amount of sunlight reaching the Earth’s surface and increasing the albedo—the proportion of radiation reflected to radiation absorbed—of large portions of newly desertified range-land.398

Bryson was not the only scientists who saw the human hand as a climatic force.

In 1971, Rasool and Schneider’s paper put pollution into the equation, with surprising

results. They incorporated aerosols from industry—SO2 the most important among them—into a simple atmospheric model, and they, like Bryson, found the cooling effect

of the aerosols more than canceled out the warming effects of CO2. Combined with smoke and dust from agricultural activities and from volcanoes, industrial pollution might

398 Reid A. Bryson and Thomas J. Murray, Climates of Hunger: Mankind and the World’s Changing Weather, (Madison: University of Wisconsin Press, 1977), 137; J. Murray Mitchell, Jr., “A Preliminary Evaluation of Atmospheric Pollution as a Cause of the Global Temperature Fluctuation of the Past Century,” in Singer, Global Effects of Environmental Pollution, 139-155; Bryson and Wendland, “Climatic Effects.”

179 not only overpower the warming effects of CO2; it might create a cooling sufficient, in the right conditions, to trigger another ice age.399

Challenged by proponents of the cooling theory, modelers at GISS, NCAR, and the Geophysical Fluid Dynamics Laboratory (GFDL) redoubled their efforts to create models that could handle an increasingly complex mix of variables. They started from one-dimensional models that essentially described vertical columns of air independent of the influences of seasonality and landscapes—models like the one Rasool and Schneider had used to demonstrate the potential climatic effects of industrial aerosols.400 Working with an influx of observed and measured data from the Global Atmospheric Research

Program (GARP), modelers at participating institutions (like NCAR and the GFDL) pushed their programs to incorporate new and different aspects of the global climate system. With every improvement in resolution beyond the static, one-dimensional models of the 1960s, new ideas cropped up about what new questions the model might help to answer. Atmospheric modelers like Rasool, Schneider, Manabe (GFDL), Joseph

Smagorinski (GFDL), Warren Washington (NCAR), Akira Kasahara (NCAR), and James

Hansen (GISS) worked to increase the capacity of their programs to deal with large and medium scale externalities and feedbacks from throughout the system. With increasing computer power, they made remarkable gains.

399 Interview with Stephen Schneider by Bob Chervin, Jan 10-13, 2002; Rasool and Schneider, “Atmospheric Carbon Dioxide and Aerosols,” 138. 400 The modeled atmosphere has four potential dimensions: longitude, latitude, elevation, and time. Taken together, these four dimensions make up a General Circulation Model (or an “ocean-atmosphere” circulation model, depending on the variables that one looks at). The 1967 Manabe-Weatherald model, on which Rasool and Schneider built their first study, presented atmospheric phenomenon in terms of elevation, and it is called a “Radiative-Convective Model.” Latitudinal one-dimensional models, produced separately by Michael Budyko and William Sellers in 1969, by contrast, describe the Earth’s energy balance in terms of variations in radiation between the equator and the poles. For more see Dennis L. Hartmann, “Global Climate Models,” Chapter 10 in Global Physical Climatology (San Diego, CA: Academic Press, 1994), 254-285.

180 These increasingly sophisticated computer models tended to reaffirm atmospheric

scientists’ initial conclusions that increasing CO2, rather than aerosols, represented the biggest threat to global climatic stability. In the new models, airborne solids presented enormous complexity. They didn’t all act to cool the Earth. As some climatologists already knew, atmospheric dust by itself both reflects and absorbs sunlight. It can also provide the nuclei for the formation of a reflective cloud layer, which in turn might, depending on the type of clouds, reduce the amount of solar radiation reaching the

Earth’s surface, producing a cooling effect.401 But the new models showed that the height, duration, and geographical distribution of aerosols also made a difference in the type of effect they could have. Some aerosols contributed to cooling, as Bryson,

Schneider, and Rasool all had originally predicted, but others actually contributed to warming.402 Schneider himself pointed out that some of his calculations in the paper with

Rasool didn’t stand up to more complex models.403 As modelers incorporated new dimensions like variable humidity, seasonality, and differences between tropospheric and stratospheric behavior into their models, the increases in carbon dioxide attendant to fossil fuel consumption once again appeared as a more powerful force than aerosols in affecting the overall radiation budget of the globe.404

401 Both Bryson and Mitchell used the uncertainties of their work and the overall dearth of research on non- volcanic atmospheric dust as a way to question the wisdom of the SST program, which, according the scientists, would put enough particulate matter in the right place to create a 5-10% increase in cirrus clouds of the Northern Hemisphere—“not negligible” in terms of climatic change. Because water vapor is adept at absorbing energy, however—and is in fact the most important and widespread greenhouse gas—clouds could actually go both ways in a warming-cooling equation, depending on their structure. Bryson and Murray, Climates of Hunger, 137. Mitchell, “A Preliminary Evaluation,” 139-155. 402 Schneider, The Genesis Strategy, 136. 403 As Schneider often notes, the limitations of his earlier model were more mathematical than theoretical. Interview with Stephen Schneider by Bob Chervin, Jan 10-13, 2002. 404 Schneider briefly described the difference between his 1971 model, based on the Manabe-Weatherald model, and what would come later in his interview with Bob Chervin in 2002. For a more technical contemporary description of the available models in 1975, see Stephen H. Schneider, “On the Carbon

181 Bryson and other climatologists still weren’t convinced, but by the late 1970s, warming seemed to many scientists the most likely climatic problem. “If man-made dust is unimportant as a major cause of climatic change,” wrote oceanographer Wally

Broecker of Columbia University’s Lamont-Doherty Geological Observatory in Science

(a contestable assertion in 1975), “then a strong case can be made that the present cooling trend will, within a decade or so, give way to a pronounced warming induced by carbon dioxide.”405

Divisions between partisans of warming and those of cooling increasingly reflected differences over methodologies, which varied between disciplines. Steeped in a tradition of measurement and empiricism, more traditional meteorologists and climatologists like Bryson and former Weather Bureau Director of Climatology Helmet

Landsberg distrusted the theoretical nature of general circulation models, which failed, in their minds, to account for the observed conditions of the climatic past. Historical meteorologists, geologists, and climatologists all worked with computer models, but they preferred physical, documentary, and statistical evidence of real past events, which gave them reference points for addressing the possible causes of a contemporary climatic shift.

General circulation modelers looking at CO2, though they disagreed emphatically over details of how to structure models and what those models might actually reveal, retorted

Dioxide-Climate Confusion,” Journal of the Atmospheric Sciences, vol. 32, issue 11 (November, 1975), pg. 2060-2066. 405 The Lamont-Doherty Geological Observatory, originally the Lamont Geological Observatory, would later go on to be named the Lamont-Doherty Earth Observatory in 1993 in recognition of the growing influence of the Earth sciences. For a very brief history of LDEO, see http://www.ldeo.columbia.edu/about-ldeo/history-lamont-0; Wallace S. Broecker, “Climatic Change: Are We on the Brink of a Pronounced Global Warming?” Science, vol. 189, no. 4201 (August 8, 1975): 460. In

November of 1975, Schneider attempted to summarize the vast uncertainties involved in CO2 modeling in a relatively technical article in the Journal of the Atmospheric Sciences, lamenting the complexities of “climatic feedback mechanisms not accounted for in state-of-the-art models.” He concurred with Broecker, and suggested that a doubling of atmospheric CO2 should lead to a 1.5-3°C increase in global temperatures. Schneider, “On the Carbon Dioxide-Climate Confusion.”

182 that the traditional standard of empiricism had to be dropped in theoretical constructions of future climates. The atmospheric conditions of those climates—twice the atmospheric

406 CO2 in this case—had no historical corollary. Instead of relying on the historical- statistical methods of geologists and other climatologists, general circulation modelers set theoretically plausible boundary conditions and then ran numerical simulations of atmospheric phenomena based on the laws of fluid mechanics.407 These models essentially described simplified interactions between components of the atmosphere, including the dust and aerosols that so concerned Bryson.

The disciplinary divisions that shaped the cooling-warming debate also tended to cut along loose institutional lines. University climatologists like Bryson, H.H. Lamb,

Thomas Murray, and Robert Matthews promoted further research (i.e. greater funding) for historical and statistical climatological studies in order to compete with atmospheric modelers at well-funded government and quasi-governmental institutions like NCAR,

NOAA’s GFDL, and NASA’s GISS, who already received the vast majority of government funding for climate science.408 These disciplinary and institutional categories

406 Schneider recalls a confrontation with Landsberg at a AAAS meeting in Baltimore in 1973 in which Landsberg attacked him on this point. Interview with Schneider by Bob Chervin, Jan 10-13, 2002. Humanists have recently begun to revisit the argument about the value of history in understanding climate change, most notably in J.R. McNeill’s “Can History Help With Global Warming,” in Climatic Cataclysm: The Foreign Policy and National Security Implications of Climate Change, edited by Kurt Campbell (Washington DC: Brookings Institution Press, 2008), 26-48. 407 In reality, both the statistical methods of historical climatologists and the numerical methods of general circulation modelers are more complex than described here, and practitioners in both general groups frequently incorporate concepts and methods from each other. To the extent that these methodologies delineated differences in the cooling-warming disagreement, they have historical significance. They do not, however, hold water in describing the current climate science community. 408 A CIA report on the state-of-the-art in climatology somewhat comically—though not inaccurately—identified three mains school of thought in climate science, the Lambian school, named for the statistical and historical work of H.H. Lamb, the Smagorinsky-ian school, relying on general circulation modeling and named for GFDL director Joseph Smagorinsky, one of the fathers of numerical weather prediction, and the Budyko-ian school, named for Soviet climatologist Michael Budyko, whose analytical work on the earth’s energy budget in the ‘50s and ‘60s, without the benefit of technological advances in computers and computer modeling, simplified the global system into a series of pencil-and-paper type

183 were relatively flexible, however, and they represented only one aspect of the internal divisions of the climate science community.

The Continuing Battle for “Good Science”: Climate Science and the Legacy of the SST

In the emerging politics of climate, it was not the direction of climatic change that mattered most, but its severity. Both the warming scenario and the cooling scenario threatened the stability of environmental, agricultural, and economic systems by increasing “climatic variability,” or the year-to-year fluctuations of local and regional weather patterns.409 Outspoken scientists like Bryson and Schneider emphasized the need to incorporate the risks of climatic instability into national and international government infrastructures, regardless of the direction of change. More conservative scientists, on the other hand—some conservative politically, some only professionally—were wary of the crisis mentality that pervaded popular scientific literature in the 1970s, and of the unspecified government action it implicitly supported. With the SST debate fresh in their minds, they feared that overzealous and premature policy recommendations absent a more definitive scientific consensus might both undermine climate scientists’ authority as experts and undercut government support for their projects. The scientific disagreement

equations describing a global energy budget. According to the CIA, the Smagorinksy-ian school was the dominant methodology within the U.S. government, and received over 90% of U.S. government science funding (for what is a little vague). The report identifies the University of Wisconsin—and Bryson and John Kutzbach in particular—as “the focal point for climatological research in the United States. “The Lambians and their primarily statistical approach are beginning to lose favor,” the report notes,” but their development of historical climatological records has provided a vital service within the climatological community.” “A Study of Climatological Research as it Pertains to Intelligence Problems,” Working Paper of the Office of Research and Development, August 1974 in The Weather Conspiracy: The Coming of the New Ice Age, report by the Impact Team (New York: Ballantine Books, 1977), 161-196. 409 Louis M. Thompson, “Weather Variability, Climatic Change, and Grain Production,” Science, vol. 188, no. 4188, Food Issue (May 9, 1975), 535-541.

184 about climate change, like the SST controversy, soon also became a debate about “good science” and its relationship to policy.

Fallout from the SST controversy helped shape discussions about the relationships between climate science and public policy in the early 1970s. The scientific controversy

over the SST primarily involved concerns about NOx and stratospheric ozone depletion, but between 1971 and 1975, research into supersonics increasingly focused on climate as well.410 After Congress essentially killed the American SST by cutting its funding in

1971, Nixon’s Department of Transportation decided to launch a comprehensive, government-sponsored study of the supersonic’s potential atmospheric impacts in order to inform U.S. policy on allowing foreign SSTs like the French Concorde to land at

American airports.411 Mustering $21 million over 3 years, the Department of

Transportation’s Climate Impacts Assessment Program (CIAP) brought a wide range of

410 Harold Johnston, a major opponent of the SST from start to finish, recalled this dual mandate in a Congressional hearing on FAA certification of supersonics in 1975—a hearing in which Johnston effectively demonstrated how the CIAP report’s executive summary had been intentionally designed to deceive its readership into believing CIAP’s scientists had essentially cleared the SST. In reality, despite the title of the study, CIAP revealed little in terms of the impacts of supersonics on climate, which took a back seat to the much more pressing public health issue of ozone reduction. Indeed, though the study raised questions about possibility of particulates from SST exhaust creating cooling stratospheric cirrus clouds, the CIAP report concluded that the climatic effects of a fleet of 500 supersonics—and certainly of the 20-30 Concordes that might ply the air in the next few years—would have a negligible effect on global climate. CIAP did, however, pay both scientific and financial dividends for scientists studying climate. For example, not only did Schneider apply for and get a grant for NCAR post-doc Jim Coakley to help build a more sophisticated model that included elements of the stratosphere, the results of his work for CIAP led him to identify the flaws in his own study on aerosols with Rasool, leading to a more developed concept of “climate sensitivity” that provided some nuance to the warming-cooling discussion. U.S. Congress, House Committee on Government Operations, Hearing, FAA Certification of the SST Concorde, 94th Congress, 1st Session, 1975 (Washington, D.C.: U.S. Government Printing Office, 1975), 67-93; Interview with Stephen Schneider by Bob Chervin, Jan 10-13, 2002. 411 A cynic might reasonably argue that the DOT hoped to used CIAP as a way to keep the SST issue alive, hoping that inconclusive science might score proponents of SST a public relations victory and get the U.S. version of the plane back on track. In any case, the Administration badly needed justification for allowing foreign SSTs to serve American destinations despite their inability to meet U.S. environmental regulations—a shortcoming readily apparent to environmentalists like Representative Richard Ottinger of New York (who was also the executive vice president of Friends of the Earth). Without clearance from the U.S. FAA, the Concorde would lose its primary routes, leaving its French and British government backers in an embarrassing lurch. U.S. Congress, FAA Certification of the SST.

185 scientists from government agencies, public and private research institutions like the

Center for the Environment and Man (part of Traveler’s Insurance Company) and NCAR, and individual universities under the financial umbrella of a single, long-term government project.

CIAP’s executive summary was equivocal about the SST. Diverging from the dominant views of its participating scientists, the Department of Transportation’s report downplayed both the Concorde’s and the Boeing 2707’s potential environmental impacts.

The CIAP group confirmed that a fleet of 500 Boeing 2707 SSTs would decrease atmospheric ozone by about 12%, but the DOT nevertheless recommended extending the

Concorde rights to land at American airports. American Geophysical Union Solar

Planetary Radiations Section President and CIAP reviewer Thomas Donahue complained in a letter to Science in March of that year about the executive summary of the report. It

“conceals the logical conclusions of the study as they were presented in

monographs…inserts new concepts concerning ultimate SST fleet sizes, flight

times, and emissions standards without candidly stating the effects on the

stratosphere of such fleets…and it, together with uncorrected stories based on it

and press interviews accompanying its release, have [sic] caused a serious loss of

credibility to atmospheric scientists.”412

“The Associated Press (AP) wire story,” Donahue continued, “had the effect of concealing, and even negating, the fact that CIAP actually supported predictions made by

412 Thomas M. Donahue and Alan J. Grobecker, “The SST and Ozone Depletion,” Science, vol. 187, no. 4182 (March 28, 1975): 1142.

186 McDonald, Crutzen, and Johnston in the early 1970s.”413 With no intention of releasing statements condemning a pet program of the high technology lobby, the DOT manipulated the message. Their meddling prompted bitter responses from many scientists.414

Few in the general public got their information from Science, however, and the media took a different view than Donahue. The mainstream press used the report—or, more accurately, the executive summary of the report—to beat cautionary scientists and environmentalists over the head. “It must, in retrospect, seem ludicrous,” The Christian

Science Monitor mused, “that Congress…should have halted the SST in mid-construction for reasons now proved to be myths—not only unproved but proved to be wrong.”415 The

New York Times treated the report similarly.416 An editorial in The San Francisco

Chronicle was less charitable, blaming “the Eco-Freaks and their allies in the Congress” for their “well-orchestrated use of the big lie technique….But now the facts are in. The anti-SST people were wrong.” “The SST and the Disaster Lobby,” the headline read.417

413 Paul Crutzen, one of Johnston’s students at Berkeley, would go on to play a major role in the fight to ban CFCs and protect the ozone. Alongside Mario Molina and F. Sherwin Roland, he won a for his efforts in 1995. Donahue and Grobecker, “The SST and Ozone Depletion,” 1142. 414 As Stephen Schneider explained in The Genesis Strategy, the CIAP report led the media to criticize environmentalists in part because of confusion over which SST they were in fact talking about. Concerns about the Boeing SST that would have been built in the U.S. were largely vindicated in the report, but the Department of Transportation, which hoped to secure airspace and landing privileges of the British-French Concorde, focused on the Concorde, not the Boeing plane. The Concord’s engine was smaller and flew lower, and had a less pronounced effect on atmosphere and climate. Schneider, The Genesis Strategy, 191- 192. 415 In “Scientists Clear SST,” The Christian Science Monitor appears to have overlooked both the real substance of the CIAP report and the historical record. The SST’s funding was cut as Nixon tried to move from the design phase to the building phase—hardly “mid-construction.” In addition, the Monitor goes on to claim that the stratospheric ozone argument provided the grounds on which to kill the project, which, as I mentioned previously, was probably not the case. Christian Science Monitor, Feb. 5, 1975; also cited in U.S. Congress, FAA Certification of the SST, 72. 416 New York Times, “World SST Fleets Said Not to Damage the Ozone Blanket,” January 21, 1975, cited in U.S. Congress, FAA Certification of the SST Concorde, 72. 417 John D. Lofton, “The SST and the Disaster Lobby,” San Francisco Chronicle, Feb. 12, 1975, cited in U.S. Congress, FAA Certification of the SST, 68.

187 The media’s rebuke of anti-SST scientists reflected their weariness of pessimistic reports on global environmental problems in the 1970s. Some scientists agreed. Even before the CIAP report, Nature editor John Maddox lamented what he called “The

Doomsday Syndrome” among environmental scientists more generally. He equated works like Commoner’s The Closing Circle and Ehrlich’s The Population Bomb with sidewalk “sandwich boards proclaiming ‘The End of the World is at Hand.’” Even

“more soberly written” investigations of the monde problematique like Ehrlich’s

Population, Resources, Environment, the Club of Rome’s The Limits to Growth, and

Roger Revelle’s 1971 The Survival Equation, offended Maddox because of their pervasive pessimism and crisis mentality.418 For many journalists, the CIAP report vindicated Maddox’s skepticism.

Some climate scientists, however, sought to emulate environmental and biological scientists’ popular exposés, and their work rekindled debates about scientists’ appropriate role in society and politics. Despite their differences, partisans on both sides of the warming-cooling divide shared a common precautionary concern for what they considered climatically risky human behaviors. Scientists like Stephen Schneider and

Bill Bryson hewed to the ideals of objectivity and neutrality in their professional work, but they also believed that “good science” could—and should—inform political decision-

418 John Maddox, The Doomsday Syndrome (New York: McGraw Hill, 1972); Barry Commoner, The Closing Circle: Nature, Man, and Technology (New York: Alfred A. Knopf, Inc., 1971), 11; Paul R. Ehrlich, The Population Bomb (New York: Ballantine Books, 1968); Donella H. Meadows, et al., The Limits to Growth (New York: Universe Books, 1972); Paul R. Ehrlich and Anne H. Ehrlich, Population, Resources, Environment: Issues in Human Ecology (San Francisco: W.H. Freeman and Company, 1970).

188 making.419 In 1976 and 1977, Schneider and Bryson each published a popular book aimed at putting the troubling links between climate and resources into the public eye.

In Climates of Hunger: Mankind and the Changing Weather, Bryson and collaborator Thomas Murray sought to expose the vulnerabilities of marginal agricultural systems to inevitable climatic variation through a series of historical analogies from ancient Greece to Early Modern Europe. The founding chair of the University of

Wisconsin’s meteorology department, Bryson had become the first director the university’s Institute for Environmental Studies in 1970.420 He and Murray saw climate and weather as driving forces of environmental and social change. “Climatic variation,” the authors warned, “like death and taxes, is certain.”421 Dismissive of the as the ultimate cause of change, the authors nevertheless warned that “climate tends to change rapidly rather than gradually,” and that “variable climate brings serious problems for world food production.”422 Humans themselves had contributed to increased climatic variability through their emissions of aerosols and dust; they had, according to the authors, been “living on favorable climate, nonrenewable energy, and other good fortune that cannot continue.”423 Climates of Hunger warned that upcoming global-scale changes in climate would reveal the drawbacks of ignoring natural systems in the course of development. Governments and societies would be left with an intractable set of resource problems that had historically led other civilizations to collapse.

419 Schneider in particular was very clear on scientists’ responsibility to inform policy, especially on the SST. “Would you prefer to estimate the SST impact,” he asked, “or have Senator Goldwater guess it for you?” Schneider, The Genesis Strategy, 191. 420 “Reid Bryson, Emeritus Professor,” http://ccr.aos.wisc.edu/bryson/bryson.html. 421 Bryson and Murray, Climates of Hunger, xii. 422 Ibid., 153, 155. 423 Ibid., 119.

189 Bryson’s book was more a popular exposé than a real proposal for action, but some of its academic reviewers addressed its shortcomings in terms of scientists’ political credibility.424 In the Annals of the Association of American Geographers, for example,

David Greenland chastised Bryson for overstating his contestable conclusions in such a public forum. “There are implications,” he wrote, “that the book is aimed at politicians in an attempt to influence public opinion.”425 Greenland worried that publicizing unresolved or contested scientific issues undermined scientists’ authority as experts and threatened their ability to garner resources and attention for their research. “If, for example,” he wrote, “the politician reading this book is already predisposed against allocating more funds for climatic research, he will have greater ammunition if he discovers climatologists are portraying hypothesis as theory.”426

Stephen Schneider’s The Genesis Strategy: Climate and Global Survival directly challenged this type of scientific conservatism. Convinced that the risks of waiting to mitigate potential catastrophe far outweighed the rewards of “ill-founded certainty,”

Schneider and his co-author Lynn Mesirow criticized the scientific community’s commitment to “consensus” on poorly understood issues with high stakes like climatic change. “Consensus,” he contended, “is a poor way to do science.”427 The book’s central

424 In the forward, Canadian climatologists F. Kenneth Hare of the University of Toronto’s Institute for Environmental Studies referred to the academic community as a “narcissistic coterie”—not the book’s intended audience. In his review of the book for the Annals of the Association of American Geographers, David Greenland acknowledged Hare’s statement without acknowledging Hare himself until later in the article, and the review implies that the words belonged to Bryson and Murray. Bryson and Murray, Climates of Hunger, ix; David Greenland, “Book Review: Climates of Hunger: Mankind and the World’s Changing Weather,” Annals of the Association of American Geographers, vol. 69, no. 2 (Jun., 1979), pp. 319-320. 425 Greenland, “Book Review: Climates of Hunger.” 426 Ibid. 427 As I will discuss in further chapters, Schneider would go on to work as a consultant and lead author on a number of WMO and IPCC reports for three decades, and continues to use the consensus reports of the IPCC not only as a way to disarm many of his critics, but as an example of how scientists have managed to make an impact in the political fight to get governments to take action on global warming. His current

190 metaphorical conceit—a comparison of climate scientists’ warnings to Joseph’s Old

Testament plea to Pharaoh to hedge against a potential famine by storing food—reflected a conscious abandonment of the “traditional role of the scientist to advance knowledge quietly” in favor of a bolder activism that mixed science and policy. 428

Schneider’s was a more ambitious book than Climates of Hunger, and a more polemical one. Schneider and Mesirow called on “our global society to build into our means of survival sufficient flexibility and reserve capacity to hedge against the repeated climatic variations that have been so well documented in history.”429 Following the example of environmental scientists and biologists like Barry Commoner and Schneider’s friend Paul Ehrlich, Schneider and Mesirow used the platform of climate science to expound more generally upon the same monde problèmatique at the heart of more comprehensive studies like The Limits to Growth. Schneider and Mesirow consciously ventured “beyond the confines of [Schneider’s] academic training.”430 “The dangers associated with climatic change,” they wrote,

“are merely a part of the entire world problèmatique, and to view them in

isolation from the rest of the world predicament would be to repeat the mistakes

positions are not necessarily incommensurate with those expressed in The Genesis Strategy. One can imagine with good humor, however, and discussion about consensus between the Stephen Schneider of 1976 and the Stephen Schneider of today. Schneider, The Genesis Strategy, pg. 136. Schneider makes a very similar comment about consensus on pg. 10. 428 Schneider , The Genesis Strategy, xiv. 429 Ibid., xvii. 430 Ehrlich, along with his wife, Ann Ehrlich, served as something of a mentor for Schneider and Mesirow on the project. The foursome exchanged frequent letters on the book, and the Ehrlichs—particularly Paul—provided Schneider and Mesirow detailed and often very harsh feedback on manuscripts. In a particularly frank missive on July 17, 1975, for example, Ehrlich criticized Schneider for “sounding pompous in [his] moralizing.” “You don’t bring to the moral analysis anything like the precision you require in your scientific work,” Ehrlich noted. He also questioned Schneider’s use of terminology from other scientific fields. “Frankly,” he wrote, “I don’t think you know your ass from first base about carrying capacity except to the extent that meteorological changes may affect agriculture.” Paul Ehrlich to Steve Schneider, July 17, 1975, made available to the author by Stephen H. Schneider, Stanford, California; Schneider , The Genesis Strategy, pg. xiv.

191 of many narrowly specialized observers who have examined the prospects for the

future only through the tunnel of their expertise.”431

In contrast to Bryson, who put his professional scientific work in a popular format in order to convey his concern over the possible societal impacts of climatic change,

Schneider used his position as a climate scientist to establish credibility in a broader analysis of the interactions between climate, food, energy, and politics. He then laced this analysis with pointed commentary and policy proscriptions that went far beyond the call for more funding traditionally acceptable within the scientific community.432

Puffed on the cover by Ehrlich, Margaret Mead, and Carl Sagan, The Genesis

Strategy reached a remarkably wide audience. The book earned Schneider reviews in

The New York Times and Washington Post, and landed him, like Ehrlich and Sagan before him, in a seat across from Johnny Carson on The Tonight Show.433

But the book also raised the ire of many within a community of atmospheric scientists already up in arms about the proper relationship of scientists to society after the

SST debate and the CIAP report. Helmet Landsberg, the director of the University of

Maryland’s Department of Atmospheric Science and Meteorology, disagreed both with

431 Emphasis in original. For more on the term monde probèmatique, see Chapter 3. See also Fernando Elichirigoity, Planet Management: Limits to Growth, Computer Simulation, and the Emergence of Global Spaces (Evanston, IL: Northwestern University Press, 1999; Schneider, The Genesis Strategy, pg. 246. 432 As Schneider notes in the acknowledgements, the bulk of the book reflected Schneider’s experiences, opinions, and expertise…as well as his personality, which, though Schneider doesn’t note it, I consider a credit to Mesirow’s skills as an author. In effect, partly because of Mesirow’s ability to channel Schneider’s voice, the book is really Schneider’s, and I treat it as such throughout. 433 Schneider made four appearances in the spring and summer of 1977 as a result of the success of the book, one with Steve Allen during Carson’s recovery from a pinched nerve in May, and three with Carson himself in June, August, and September. The appearances arose in large part thanks (according to Schneider) to recommendations from Ehrlich and Sagan, although he was not invited back after he deviated from the script of the fourth show. He also made an appearance on the final episode of Dan Rather’s CBS special “Who’s Who,” in which he and Rather were filmed fishing in Barker Dam Reservoir in Nederland, Colorado (near Boulder). John Carmody, “The TV Column,” The Washington Post, May 26, 1977; Don Shirley, “The TV Column,” The Washington Post, July 23, 1977; Interview with Stephen Schneider by Bob Chervin, Jan 10-13, 2002.

192 Schneider’s science and with his popular approach. He was particularly critical of

Schneider’s “use of climatic models for predictions to guide public policy.”434 Quoting from widely respected GFDL global circulation modeler Joseph Smagorinski, Landsberg reminded Schneider that “‘crude or premature estimates can be very misleading in providing guidance or such far-reaching decisions and may be far more damaging than no estimate at all.’”435 He emphasized Smagorinski’s warning that “‘we should be wary of basing broad national or international decisions on hand-waving arguments or back-of- the envelope calculations.’”436

Politically conservative scientists like Landsberg primarily attacked Schneider for prematurely and recklessly venturing into policy and popular culture with highly uncertain scientific conclusions, but they also objected to his liberal policy proposals. In what F. Kenneth Hare called “devastatingly naïve” policy recommendations, Schneider advocated an array of liberal measures that would expand the size and power of the federal government.437 He semi-seriously proposed a fourth branch of government, the

“Truth and Consequences Branch,” a scientific body meant to study and provide information on the long-term impacts of current policies and actions. He also called for a government-funded U.S. grain reserve, various global environmental treaties, an

434 H.E. Landsberg, “Forum,” E.O.S.: Transactions of the American Geophysical Union, vol. 58, no. 3 (March, 1977): 122. See also Landsberg, “Review: The Genesis Strategy—Climate and Global Survival,” E.O.S.: Transactions of the American Geophysical Union, vol. 57, no. 9 (September 1976): 634-5. 435 Ibid. Landsberg cites Joseph Smagorinski, “Global Atmospheric Modeling and the Numerical Simulation of Climate,” in Weather and Climate Modification, edited by Wilmot Hess (New York: Wiley, 1974). 436 Landsberg, “Forum,” 122. 437 Hare and Schneider corresponded amicably about the book, and as a director of an environmental science institution, Hare genuinely agreed with and liked many of Schneider’s ideas. He also agreed with his politics, however naïve. “I hope that they [Schneider and Mesirow] will not be discouraged by the braying laughter that this book may engender in such quarters [academia],” he wrote, because “real politics, not the chitchat of campuses, depends on the harnessing of inspired naivety.” F. Kenneth Hare, “Book Review,” Bulletin of the American Meteorological Society, vol. 58, no. 8 (August, 1976): 1016.

193 international inventory and collective control of nuclear materials, and a more equitable distribution of development aid and technology between the first and third world. In short, he introduced a new national and international order based on science and reason, but one also implicitly infused with Schneider’s own internationalist liberal values.

“This is a risky game for a younger man to play,” cautioned Hare in his otherwise laudatory review of the book in Bulletin of the American Meteorological Society,

“and he takes his professional life in his hands when he does so. For there is a

widely held view that scientists ought to stay out of politics…We are a

conservative profession, and I would hazard a guess that dynamicists range

themselves on our far right. So Schneider will not lack for criticism and even

abuse for having written this book.”438

As Hare and other sympathetic colleagues feared it would, Schneider’s book drew criticism from higher-ups at NCAR who, according to Schneider, didn’t find his style appropriate for an NCAR research scientist.439 His appearances on television, and particularly on Johnny Carson’s show, further stoked resentment. Adversaries at NCAR later tried to deny his bid for Senior Scientist based, at least in part, on their distaste for his popular science.440 Schneider maintained his favored status among well-respected and

438 By “dynamicist,” Hare referred to a school of meteorology and climatology that based its predictions on theoretical models that described the atmosphere in terms of the laws of fluid- and thermodynamics. Climate modelers of all sorts were—and still are—essentially dynamicists to some extent; that is, they use their understanding of physical cause and effect to predict future conditions from some set of baseline data. Hare’s point here is that Schneider has not only challenged the relationship between science and politics, but he has done so from within a particularly conservative scientific sub-field. Hare, “Book Review,” 1016. 439 Interview with Stephen Schneider by Bob Chervin, Jan 10-13, 2002. 440 Landsberg’s criticisms of the book ran the gambit, and even Hare, who’s Bulletin of the American Meteorological Society review painted the book favorably, had to agree with a few of Landsberg’s points. Landsberg reviewed the Genesis Strategy in EOS, Transactions of the American Geophysical Union, TK! Vol. ?, No. ?, September, 1976, 122. Schneider provided a rebuttal to the review, to which Landsberg was allowed to reply, and the testy exchange appeared in EOS, Transactions of the American Geophysical Union, vol. 58, no. 3, March, 1977; Interview with Stephen Schneider by Bob Chervin, Jan 10-13, 2002.

194 progressively-minded Senior Scientists at NCAR like Kellogg and Roberts, but as a colleague recalls, “even at NCAR, Steve was pretty isolated for a while because of that book.”441

Climate, Food, and “The Environment”

Internal disputes over “good science” and the limits of scientific advocacy did not play out in a vacuum. Indeed, climate scientists’ concerns over global climate change developed in tandem with an international environmental movement focused on environmental degradation at a global level, and it was this international movement that provided the primary context for scientists’ forays into environmental advocacy. In the late 1960s and early 1970s, scientists who studied elements of climate and weather from various disciplines sought to apply their basic research to a global environmental crisis that included problems of energy use, agricultural waste, industrial pollution, and ecological change. These scientists recognized that the processes of climate and climatic change entwined with environmental degradation at every level, and they began to work together to incorporate climate into large-scale environmental assessments. The SCEP and SMIC studies contended that humans could inadvertently modify regional and global climates through deforestation, overgrazing and marginal-lands agricultural development,

industrial air pollution, and the release of trace gases like CO2 and NOx into the atmosphere. The resulting changes had the potential to affect natural and human environments already threatened by problems associated with population growth, consumption, and technological development. It was these types of practical issues—or,

441 Michael Glantz, in discussion with the author, National Center for Atmospheric Research, Boulder, CO, 11/12/07; Interview with Stephen Schneider by Bob Chervin, Jan 10-13, 2002.

195 rather, a subset of these practical issues related to climate—that most concerned climate scientists in the 1970s.442

At the time, the most important practical problem related to climate and climatic change was food supply. Over the course of the 1960s, the world’s population had continued to grow at an alarming rate, but so too did technological developments in agriculture. This “green revolution,” combined with generally favorable weather, made feeding new mouths possible.443 In the late 1960s and early 1970s, however, a series of climatic and meteorological anomalies led to an alarming set of shortages in the world food supply. In the African Sahel region, four consecutive years of drought produced major collapses in the regional food supply. The shortages resulted in large-scale migrations, social unrest, and starvation. In 1972, the Indian monsoon came a week late, stressing a food infrastructure already operating on a thin margin. A drought in the

Soviet Union also caused major food shortages, leading to mutually embarrassing purchases of 18 million tons of grain from the United States, where domestic food prices began to rise. For a brief period in late 1973—just before the Christmas Eve oil price shock of that year—a bushel of wheat sold for twice the price of a barrel of oil.444 As

442 Of the 26 principles outlined in the Conference’s “Declaration on the Human Environment,” seven involved promoting development. “Declaration of the United Nations Conference on the Human Environment,” United Nations Environment Programme, www.unep.org/Documents.Multilingual/Default.asp?DocumentID=97&ArticleID=1503 ; see also New York Times, “Text of the Environmental Principles,” June 17, 1972. 443 See Nick Cullather, “Miracles of Modernization: The Green Revolution and the Apotheosis of Technology,” Diplomatic History, vol. 28, no. 2 (April 2004), 227-54. See also Vandana Shiva, The Violence of Green Revolution: Third World Agriculture, Ecology, and Politics (London: Atlantic Highlands, 1991). 444 For a longer, more detailed (and entertaining) popular account of the controversial U.S. grain sale to the Soviet Union, see James Trager, The Great Grain Robbery (New York: Ballantine Books, 1975), which is an updated version of Trager’s Amber Waves of Grain: The Secret Russian Wheat Sales That Sent American Food Prices Soaring (New York: Arthur Fields Books, 1973). See also Lester Brown with Erik P. Eckholm, By Bread Alone, (New York: Praeger, 1974), 5; John C. Whitaker, Striking a Balance: Environment and Natural Resources Policy in the Nixon-Ford Years (Washington, D.C.: American Enterprise Institute for Public Policy Research, 1976).

196 Lester Brown of the Overseas Development Council estimated in 1974, climatic variation led to an overall reduction of the world food supply of about 1% in 1972—the first reduction of its kind in over a decade.445 “No single factor has a greater impact on food production in any country than weather,” Brown wrote. “The vulnerability of the supply- demand balance to the weather suggests that the climate itself might well replace pollution as the dominant global environmental concern.”446

Because of its relationship to food, climate change became at once a scientific, environmental, and geopolitical concern in the 1970s. The U.S. provided three quarters of the world’s grain exports in the 1970s, and with the U.S. grain stocks dwindling and acreage previously held out of production in the U.S. land bank increasingly pressed into service, the CIA expressed concerns over the international political consequences of another poor harvest.447 “The stability of most nations is based upon a dependable source

445 Brown, By Bread Alone, 60; Brown is cited in Schneider, The Genesis Strategy, 7. 446 For historians, the connection between food and climate has been well established. As early as the 1949, Fernand Braudel built a picture of history “from the ground up,” beginning with regional environmental constants like geography and climate, which in turn helped to determine socially and culturally significant factors like food consumption and long-term patterns of trade. After outlining the bulk of the book by 1939, Braudel, the story goes, wrote his great structuralist work from memory in a Nazi prison during the Second World War. The original French version of The Mediterranean and the Mediterranean World in the Age of Philip II, trans. by Siân Reynolds (New York: Harper and Row, 1972), was published in 1949. Taking Braudel’s structuralist approach to European History a step further, Emmanuel Le Roy Ladurie made climate the subject of a thousand year historical survey in 1967, Times of Feast, Times of Famine: A History of Climate Since the Year 1000, trans. by Barbara Bray (New York: Noonday Press, 1971). Ladurie focused primarily on the historical impact of changes in climate on the European food supply. “Climatic history,” he argued, constituted part of the study of “ecological history, asking such questions as whether the fluctuations of climate—or to put it more modestly, the brief fluctuations of meteorology—have reacted on the human habitat; on harvests and thus on economy; on epidemics and diseases, and thus on demography.” Ladurie hoped throughout the 1970s that historians would play a more prominent role in studying climatic change, using their specific archival skills to corroborate physical evidence from the early modern period with physical evidence gathered from tree rings, peat bogs, and ice cores. He explained the historian’s role in an essay, “History Without People” in The Territory of the Historian, trans. by Ben and Siân Reynolds (Chicago: University of Chicago Press, 1979). See also Brown, By Bread Alone, 4, 67. 447 Two internal CIA reports, “A Study of Climatological Research as it Pertains to Intelligence Problems,” Working Paper of the Office of Research and Development, August 1974, and “Potential Implications of Trends in World Population, Food Production, and Climate,” Central Intelligence Agency, Directorate of Intelligence, Office of Political Research, OPR-401, August 1974, were made public as official documents in May of 1976. They were published as appendices in an alarmist popularization

197 of food,” read a CIA working paper on “Climatological Research as it Pertains to

Intelligence Problems,” “but this stability will not be possible under the new climatic era.”448 By 1974 the Agency “had obtained sufficient evidence” of a global cooling trend to put climatic change on the geopolitical radar. “Climate is now a critical factor,” the

CIA confirmed. “The politics of food will become the central issue of every government.”449

The relationship between climate and food also prompted some institutions to expand the disciplinary reaches of their climatic research. NCAR’s Mickey Glantz embodied this disciplinary expansion. As a former Ford Motor Company engineer and liberal professor of political science, Michael Glantz—“Mickey” to anyone he’d met—was an unlikely climatologist. Teaching at Lafayette College with a Ph.D. from the University of Pennsylvania, Glantz was “into revolutions…Yemen, Cuba, the Congo, whatever,” and had a particular fascination with Africa. Later, as a Senior Scientist at

NCAR, he remembered that he “wasn’t really a Marxist…but let’s say a realist…a bleeding heart liberal.”450 After a 1972 seminar at Syracuse University on “Drought in

West Africa,” Glantz applied for a post-doc at NCAR, and largely because of the appeal

written by a group of “investigative journalists” calling themselves the “Impact Team” in 1977. The Weather Conspiracy, 161-196, 197-224. CIA grain export estimates appear on pg. 208. See also Deborah Shapley, “The Genesis Strategy (Review),” The New York Times, July 18, 1976. 448 “A Study of Climatological Research,” in The Weather Conspiracy, 163. 449 Ibid., 166. 450 Glantz, who describes himself as something of a scientific and political Forest Gump, had nursed an interest in Africa for more than a decade before heading to NCAR. In 1963, as a Master’s student in political science at the University of Pennsylvania (where he had already received a degree in engineering), Glantz had organized a trip to Portuguese Angola, where he traipsed around the countryside with high- ranking government and military officials during the initial stirrings of a revolution that would lead to 27 years of civil war. His experience in Angola, coincidentally, led him to a job in the Senatorial office of Claiborne Pell, whose father had served as U.S. Ambassador to Portugal, and who represented Glantz’s home state of Rhode Island. It was purely by coincidence, however, that the former employee of the Senator sponsoring a ban on military weather modification a decade later found a position working on the environmental and social impacts of the largest weather modification experiment in the world. Michael Glantz, in discussion with the author by telephone from NCAR, Boulder, CO, April 14, 2008.

198 of former director Walt Roberts’s progressive scientific vision, he decided to take a position with NCAR’s Environmental and Social Impacts Group (ESIG)—a part of the

Northeastern Colorado Hail Experiment—in 1973.451 But if Glantz was an unlikely climatologist, he was an even more unlikely hail researcher. When Roberts, using money from the International Federation of Institutes for Advanced Studies (IFIAS), secured a post-doc for him in NCAR’s Advanced Studies Program for him after a year with ESIG,

Glantz jumped at the opportunity.452

Though no longer NCAR’s director, Roberts continued to push for the socially relevant study of climate at NCAR and elsewhere in 1974, and he looked to Glantz to help him venture beyond the confines of atmospheric science. With money from the

Rockefeller Foundation and support from IFIAS, he founded the Program on Food,

Climate, and the World’s Future at the Aspen Institute for Humanistic Studies. Working with the Stockholm-based institution, Roberts hoped to incorporate climate and climatic change into the larger picture of the monde problématique for the Club of Rome’s Limits to Growth sequel in 1974, Mankind at the Turning Point. He emphasized the disastrous effects of climatic “accidents” on food supplies in the context of exponentially increasing populations and a decreasing margin of food security.453 Still concerned about the overall cooling trend that appeared to have dominated the period from 1940 to 1970, Roberts and the Club of Rome sought “multidisciplinary” research that might provide a better

451 Ibid. 452 Glantz recalls sitting in the NCAR cafeteria with Schneider, Kellogg, and ASP director Peter Gilman discussing the possibility of working on the National Hail Research Project in 1973. “I’m sitting there saying, ‘yes, I’ve always been interested in hail,’” he remembers, “while in my head I’m spelling it h-a-l-e. And then later they want me to do forecasts, and I don’t know forecasts from schmorcasts, but what the hell, why not?” Ibid. 453 Jacques Freymond, “New Dimensions in International Relations,” The Review of Politics, vol. 37, no. 4 (October, 1975), 464-478.

199 understanding not only of the problem of climatic change itself, but of possible strategies for coping with climatic change’s social and economic impacts.454

Roberts asked Glantz open-ended questions about the value of long-range climate forecasting. First, he asked, given no real world political or economic constraints, how might governments and populations use a one-year warning of climatic anomalies to avoid food, water, and other natural resource crises like those in Africa or the Soviet

Union? Secondly, how might these types of climate forecasts be used in practice, and how valuable might they be?455

Glantz took as his example sub-Saharan Africa, the arid “Sahel” region so devastated by the drought years of 1968-1972. As a political scientist, Glantz had seen conditions in West Africa on the ground. Forecasting, he understood, was by itself almost useless in Africa. The processes of transporting and storing food, selectively culling livestock herds, and securing wide food distribution networks—all tactics for mitigating a forewarned food shortage—required a strong state with legitimate and well- developed machinery for running a centralized bureaucracy, a well-developed transportation network, and sufficient initial capital. All were in short supply in the developing world. “The Sahelian states are confronted by what might be termed a hydra- headed crisis,” he wrote in the Bulletin of the American Meteorological Society in 1977.

The hydra’s heads were

“1) technical and institutional constraints; 2) legitimacy, distribution, integration,

and penetration crises; and 3) embryonic political institutions carrying the burdens

of welfare and development and the implications of rapid ‘transition from labor

454 Ibid. 455 Glantz telephone interview, April 14 2008.

200 intensive to automated industrial systems,’ ‘the weakness and fragility of the state

machinery,’ and ‘the perennial shortage of state revenue,’ among others.”456

Echoing the conclusions of the U.N. Conference on the Human Environment in

1972, Glantz demonstrated that anomalies of weather and climate could strain the already marginal agricultural—as well as political and economic—systems of developing nations.

Mitigating such environmental disasters required a focus on developing stable basic social, political, and economic infrastructures before scientific advancements like long- range forecasting could benefit the region’s populace.457

The focus on resources and infrastructure did not stop at food. A colleague and friend of Glantz at NCAR, Stephen Schneider argued that the broader problems of climate, food, and development infrastructure also included an even thornier issue: energy. Schneider framed the “food/population/technology/climate problem” as an issue of competing interests in international development, with energy use at its core.458

Industrialized nations required vast amounts of energy to grow their economies, and their ever-rising demand increased prices. Rising energy prices fueled scarcity and increased prices for the petroleum-based fertilizer necessary for growing food on the environmental

margins in developing countries. Increasing the amount of CO2 in the atmosphere, meanwhile—a product primarily of First World industry—might contribute to the very climatic anomalies that made intensive fertilizer use necessary on marginal lands in

456 Michael Glantz, “The Value of a Long Range Weather Forecast for the West African Sahel,” Bulletin of the American Meteorological Society, vol. 58, no.2 (February 1977), 150-158. 457 Facing criticism that his case study revealed more about the particularities of Africa than about the general value of long range forecasting, Glantz took as his next case study the spring wheat crop in the well-developed nation of Canada. For every day lost in harvest and distribution to weather anomalies like snowstorms or mud at either end of the harvest season, he found, eight to ten days were lost in labor disputes. Michael H. Glantz, “Saskatchewan Spring Wheat Production 1974: A Preliminary Assessment of a Reliable Long-range Forecast,” Environment Canada Climatological Studies, No. 33 (Downsview, Ontario: Environment Canada, 1979). Glantz telephone interview, April 14, 2008. 458 Schneider, The Genesis Strategy, 17.

201 developing countries in the first place.459 “The atmosphere is a resource shared by all peoples and places on earth,” Schneider wrote. “Not surprisingly, there are many political implications attached to the sharing of such a finite, continuous media by the competitive interests we call nation-states.”460

Glantz and Schneider both advocated incorporating social and political science into discussions about climate change, and they stood at the far end—and far left—of an increasingly interdisciplinary spectrum of climate scientists. Glantz’ methods and in his conclusions (not to mention his self-described “Bozo the Clown” appearance) set him and his work apart.461 But his and Scheider’s focus on resources and resource management as the central “environmental” problems associated with climate underscored the concerns of the broader climate science community. “In man’s quest to utilize global resources, and to produce an adequate supply of food,” stated a group of Quaternary geologists who met at Brown University to discuss the potential impacts of rapid shifts in weather and climate, “global climatic change constitutes a first order environmental hazard which must be thoroughly understood well in advance of deteriorating climate.”462

459 Riffing off of the I = PAT equation made famous by Paul Ehrlich that related population, affluence, and technology to overall environmental impact, Schneider contended that “the climatic disruption from energy consumption is proportional to the total amount of energy used, which is equal to the population size times the per capita energy-consumption level.” Schneider, The Genesis Strategy, 293. 460 Schneider, The Genesis Strategy, 118. 461 Glantz still takes some good-natured pride in his rebellious and sometimes ridiculous appearance in the 1970s. He fondly recalls meeting the head of the Soviet Academy of Sciences as a young scientist at formal dinner party “looking for all the world like Bozo the Clown, but without the nose.” During the first writing of this chapter, amusing NCAR portraits of Glantz from 1974 and 1980 could be found on his NCAR website. Glantz’s relationship with NCAR has recently soured, however. After more than 30 years, the financially strapped institution abruptly fired Glantz—a Senior Scientist—and dissolved his Center for Capacity Building in the fall of 2008. Many in the climate science community were outraged, and though Glantz is hardly one of the most renowned scientists involved in climatic research, his firing prompted angry letters and newspaper articles both in Boulder and in the national media. For example, Andrew Revkin, “Climate-Change Program to Aid Poor Nations is Shut,” New York Times, August 7, 2008. 462 G.J. Kukla and R. K. Matthews, “When Will the Present Interglacial End?” Science, vol. 178, no. 4057 (October 13, 1972): 191. Weart refers to the “More Money Should Be Spent on Research” conclusion repeatedly in The Discovery of Global Warming. And rightly so. As he demonstrates, scientists frequently

202 Climate Change and American Environmentalism

American environmentalists also took a keen interest in global environmental issues in the 1970s, but they responded slowly and cautiously to the issue of climate change. The U.N. Conference on the Human Environment revealed deep concerns within

America’s non-governmental environmental organizations over global population, land use, species protection, ocean dumping, toxic wastes, and deforestation, but these concerns did not extend to climate change in a meaningful way. The Sierra Club, Friends of the Earth, The National Resources Defense Council, and the National Wildlife

Federation all lobbied Congress vigorously in favor of extending and expanding the

United States’ ever-shrinking contributions to the United Nations Environment

Programme throughout the ‘70s and early ‘80s, and many of these organizations created committees or sections to deal exclusively with international issues.463 In general, however, American environmental groups continued to focus primarily on the local issues that interested their largely middle-class memberships, and they carefully selected the international issues they thought their constituencies would support. Professional environmentalists sought tangible environmental problems with definitive solutions that

used alarming facts and discoveries not as a way to generate political support for environmental causes so much as to generate financial support for further research into specific scientific problems, often within their own discipline and/or institution. In this particular case, Weart notes that some members of the Brown group actually wrote to President Nixon, “calling on the government to support intensified studies,” an example, in Weart’s view of a “general movement during the 1970s” of scientifically trained people “making contact with policy elites to address the planet’s environmental future.” Weart does not elaborate on this movement until he gets to the 1980s, however. Weart, Discovery, 81. 463 Letter, “Dear Senator,” June 13, 1973, United Nations Environment Fund 1973, Sierra Club International Program Records (4:31); Memorandum, Judith Campbell Bird to Citizen Activists, “United Nations Environmental Programme Threatened by Proposed Cut in U.S. Contribution,” April 18, 1979, United Nations Environment Program Funding, Sierra Club International Program Records (5:1); Jacob Scherr (NRDC) to Patricia Scharlin (SC), April 9, 1979, United Nations Environment Program Funding, Sierra Club International Program Records (5:1), Sierra Club Archives, Bancroft Library Special Collections, University of California, Berkeley.

203 appealed to the interests and values of their grassroots constituencies, and in the 1970s, the issue of climate change did not yet fit the bill.

The Sierra Club’s International Committee served as a central node of the

American environmental effort abroad. “No organization did more to prepare the way for the Stockholm Conference and the creation of the United Nations Environment

Programme than the Sierra Club,” wrote UNEP chief Maurice Strong to Sierra Club

International Program director Patricia Rambach (later Scharlin) upon leaving his post at the U.N. in 1975, “and none have been more effective or consistent in their support for our international activities during the crucial formative period.”464 Scharlin and her colleagues collaborated with NRDC, FoE, Zero Population Growth, the National Wildlife

Federation, and others to influence international environmental policy both within official

U.S. government delegations and as an independent supranational bloc at a series of

United Nations conferences on food, water, desertification, human settlements, and population throughout the decade.

The Sierra Club’s commitment to population control helped drive its International

Program. The Club had helped to publish and promote Paul Ehrlich’s The Population

Bomb in 1968, as well as follow-up articles by Ehrlich and his wife Anne in the early

1970s, and when the International Committee began to reevaluate its mission in 1976, members made it clear that population represented their top international priority.465

Tying over-population to over-consumption, many members saw population

464 Rambach married during her tenure as the head of the Sierra Club’s International Committee, and her name appears alternatively as Patricia Rambach and Patricia Scharlin throughout the Sierra Club International Committee records. To avoid confusion, I will henceforth refer to by her married name, Scharlin, in the main text, and will only differentiate here in the footnotes. Letter, Maurice F. Strong to Patricia Rambach, November 14, 1975, Operational Records, International Committee, 1972-1983, Sierra Club International Program Records (3:16). 465 “Sierra Club International Committee Questionnaire—Five Year Plan,” (no date), Five Year Plan, 1976- 77, Operational Records, Sierra Club International Program Records (3:10).

204 control—both in the developing and more developed countries—as a sine que non of environmental protection. It was an overriding issue of global proportions that underpinned environmental issues of all scales. “Most of our environmental problems are the direct result of over-population,” one Sierra Club chapter officer wrote. 466 “The root cause of most problems,” another claimed.467 “This seems to me a basic issue which must be solved or all will be lost!”

Only slightly behind population on the list of international priorities sat a broad set of concerns over what the Sierra Club called the “global commons”—those areas shared by many nations but technically owned by none, consisting mostly, in the Sierra

Club’s view, of ocean environments. Distinguishing issues as “international” based largely on the type of governmental or supranational agency required to deal with each particular problem, members worried that the pollution and degradation of the oceans and the loss of marine environments—along with air pollution and environmental degradation in Antarctica—might affect human and natural systems on a variety of scales. They encouraged the Sierra Club to lobby in these environments’ defense in an international forum.468 In a 1976 survey of Sierra Club chapter and group officers, well over half of the respondents identified these same “global commons” issues as the primary international concerns of their local chapters, beating out more geographically specific problems like deforestation and nuclear proliferation.469 Since, after population, concerns about the health of the Earth’s oceans and air dominated the Sierra Club membership’s

466 Emphasis in original. Ibid. 467 Ibid. 468 Ibid. 469 Deforestation did pique the concern of many members, however, and the Sierra Club International Program’s leaders considered deforestation, along with the protection of species, among their highest priorities.

205 international program in the late 1970s, it stood to reason that the Sierra Club and its associates from the environmental community would be deeply involved in the mounting national and international discussions about climate change.

Few of the organization’s chapter officials advocated a wholesale commitment to the international arena for any cause, however, let alone one as nebulous and uncertain as climate change. Roughly 20% of the members polled in 1976 recommended that the

Club commit 10% or more of its overall budget to international issues, but an almost equal percentage thought the Club should reduce its international budget to less than 3% of the total, and just less than a third hoped to keep international activities to under 5% of

470 the budget. Though scientists and members of Congress took interest in the CO2 problem early in the decade, the issue of climate change had not sufficiently piqued Club member’s concern in 1977 for the International Committee to include climate as a major part of their next “five-year plan.” Nearly a third of the 1976 survey’s respondents identified “global commons” issues as their top priority for the International Program, but the sub-heading of “air pollution and climatological studies” drew no votes as a top priority, and less than a third of the respondents identified climate as a priority at all.471

As late as 1979, the year of the WMO’s World Climate Conference, discussions of climate and climatic change within the Sierra Club remained brief, tentative, and non- committal.

The problem with making climate a Sierra Club issue, as Scharlin saw it, was that members didn’t see climate change as a “‘back yard’ kind of issue, or one that commands

470 Ibid. 471 “Table 1: Question 1A—Rank in Priority of Importance the 6 International Environmental Issues You Wish the Club Should Tackle,” from “Sierra Club International Committee Questionnaire—Five Year Plan,” (no date), Five Year Plan, 1976-77, Operational Records, Sierra Club International Program Records (3:10).

206 a short term horizon.”472 Despite a strong rhetorical and administrative commitment to international issues, the Sierra Club remained an essentially local, grassroots institution.

Scharlin noted that climate change “inspires emotional interest, has a global, international dimension, and very serious, long-term complications,” but the issue lacked the immediacy of problems like population growth or industrial air pollution that could easily be framed in terms of their local and regional impacts.473 An increasing population, for example, not only strained the world’s natural resources in the abstract; it also put pressure on America’s wilderness areas and open spaces by increasing the demand for the resources these areas contained, as well as for the wilderness experience itself.474 The problem of population growth was global, but its most important environmental impacts were local. As Scharlin pointed out, neither scientists nor environmentalists had yet managed to identify the specific threats that climate change might pose to local and regional environmental amenities.475 As a result, Sierra Club members didn’t identify with the issue, and climate change failed to show up either in the International Program’s

472 “Earthcare Center,” Conversations on Earthcare Center Information Center/Network 8/15-16/1979, Program Goals, Operational Records, Sierra Club International Program Records (3:14). 473 Ibid. 474 As Roderick Nash explains in the fourth edition of his 1967 Wilderness and the American Mind (New Haven: Yale University Press, 2001), many American environmentalists saw wilderness areas in foreign nations as an extension of this larger “back yard.” To the extent that international development threatened pristine natural areas and the animals that inhabited them, that development came into conflict with some environmentalists’ key interests. Thomas Robertson argues that Americans’ and Europeans’ attempts to protect wilderness landscapes in less developed countries—mostly in Africa—amounted to a form of environmental imperialism, wherein first world environmentalists sought to protect third world landscapes from their own inhabitants. In any case, environmentalists’ interest in international wilderness areas helps to explain the philosophical links between international environmental protection and the domestic efforts of organizations like the Sierra Club and Wilderness Society. It would take another decade, however, before the threat of climate—a threat not rooted in any particular sense of place—would be seen as a tangible threat to these extensions of the American wilderness “back yard.” Robertson, “‘This is the American Earth’: American Empire, the Cold War, and American Environmentalism,” Diplomatic History, vol. 32, no. 24 (September, 2008), 561-584. 475 Sierra Club, “Earthcare Center.”

207 flagship “Earthcare Center” goals or in the larger Club priorities for the upcoming decade.476

A 1982 set of “Criteria for International Campaigns” compiled by Executive

Director Michael McCloskey reveals just how antithetical the issue of climate change was to the Sierra Club’s framework for environmental activism in the early years of national concern over global warming. First and foremost, McCloskey argued, an international environmental goal must be achievable “within a reasonable time frame.”477

The goal must be “clear and discrete; not ill-defined, nor covers [sic] impossibly large situations,” and must deal “with conditions for which a legal or regulatory solution is possible.”478 McCloskey demanded that the issue fit neatly into some institutional forum that the Sierra Club had the competence and capability to influence, and that it strike the

Club’s members as a clear threat to their “deeply felt values.”479 The source of the threat, he argued, must be something that can be pinpointed, and the larger problem put into a real life context accessible to concerned members. “[It] can’t be too exotic nor overly technical,” he cautioned.480 As a nebulous, highly technical, scientifically uncertain, long-term global issue with little material tangibility and no real framework for suggested

solutions—of which few existed, even among scientists—CO2-induced climate change could not have been farther from McCloskey’s concept of an ideal international environmental issue.

476 Ibid; “Sierra Club Priorities: Board Sets 1980 Conservation Priorities,” International Committee, Operational Records, Sierra Club International Program Records (3:18). 477 Emphasis in Original. Michael McCloskey, “Criteria for International Campaigns,” December 1, 1982. International Committee, 1972-1983, Meetings and Conferences, Operational Files, Sierra Club International Program Records [3:19]. 478 Ibid. 479 Ibid. 480 Ibid.

208 Climate as a Resource: The National Climate Program and the AAAS-DoE Workshops

The scale and complexity that made climate change such a difficult issue for environmentalists to engage with were the very qualities that appealed to many climate scientists. To them, the uncertainties of climate change made it all the more important to study and identify its potential impacts on natural resources, both at home and abroad.

Within the scientific community, efforts to incorporate climate science into resource management took two forms. First, most climate scientists, regardless of their politics, supported a new national-level bureaucracy for coordinating climatic research and incorporating climate data into domestic agricultural planning and natural resource management strategies, the National Climate Program. The National Climate Program not only supported the needs of American farmers, ranchers, and emergency management professionals, it also established a secure role for climate scientists in formulating national resources policies, both within federal agencies and in Congress. Secondly, scientists at the American Association for the Advancement of Science hoped to extend these practical domestic applications of climate science to international development.

The AAAS made climate change a central component of its international advocacy

program, and worked to establish a firmer scientific consensus on CO2-induced warming to buttress potential future policy recommendations. For advocates of both the National

Climate Program and the AAAS initiative, the primary immediate concern was to facilitate more and better research into climate. Better knowledge, climate scientists believed, would shape good policy.

In the mid-1970s, the National Academy of Sciences released three reports on climate and resources that emphasized the practical applications of climate science. The

209 Academy’s 1975 Understanding Climatic Change: A Program for Action called for greater integration of government-sponsored climate research across disciplines and institutions in order to foster a better understanding of atmospheric phenomena and to provide useful data to policy-makers and planners in dealing with issues like agriculture and emergency management where climate played an important role.481 The Academy set up a Committee on Climate and Weather Fluctuations and Agricultural Production, which released a report on Climate & Food: Climatic Fluctuation and U.S. Agricultural

Production in 1976.482 In addition, an NAS Panel on Water and Climate within the

Geophysics Study Committee conducted a study on Climate, Climatic Change, and the

Water Supply in 1977.483 The reports contained specific technical and research recommendations, and all three recommended better coordination of climate research in order to facilitate its practical incorporation into local, state, and national level planning for agricultural and other natural resource policies.

The Academy studies were prompted in part by the concerns of members of

Congress, who also supported better coordination of climate research in the service of natural resource planning. Congressional Democrats saw the increasingly frequent and

481 Cited on Spenser Weart’s website as GARP (National Academy of Sciences, United States Committee for the Global Atmospheric Research Program), Understanding Climatic Change: A Program for Action, (Washington, DC; Detroit, MI: National Academy of Sciences, Grand River Books, 1975); Weart, The Discovery of Global Warming: A Hypertext History of how Scientists Came to (Partly)Understand What People are Doing to Cause Climate Change, www.aip.org/history/climate/index.html. 482 National Research Council, Committee on Climate and Weather Fluctuations and Agricultural Production, Board on Agricultural and Renewable Resources, Climate & Food: Climatic Fluctuations and U.S. Agricultural Production (Washington, D.C.: National Academy of Sciences, 1976). 483 Schneider, it should be noted, participated in the Panel on Water and Climate, coauthoring the first section of the report on “Water Supply and the Future Climate.” National Academy of Sciences, Committee on Climate and Weather Fluctuations and Agricultural Production, Board on Agricultural and Renewable Resources, Commission on Natural Resources, National Research Council, Climate & Food: Climatic Fluctuations and U.S. Agricultural Production (Washington, D.C.: National Academy of Sciences, 1976); National Academy of Sciences, Panel on Water and Climate, Geophysics Study Committee, Geophysics Research Board, Assembly of Mathematical and Physical Sciences, National Research Council, Climate, Climatic Change, and Water Supply (Washington, D.C.: National Academy of Sciences, 1977).

210 dire—and often contradictory—reports and predictions from climate scientists, along with an admitted lack of coordination in the national climatic research effort, as an opportunity to take the initiative on federal science policy and undercut executive control over government science agencies.484 Led in the House by California Representative

George Brown, they noted that “the present research and data-gathering in the Federal

Government is scattered, fragmentary, and inadequate compared to the impact of ignorance in this area.”485 Brown and his colleagues in the Senate pushed for a

Congressionally legislated National Climate Program that would establish the framework for national-level coordination of climatic research, monitoring, and warning systems.486

In 1976, they introduced the bi-partisan National Climate Program Act. The Act’s supporters accepted that “longer term changes in climate, whether occurring naturally or resulting from human activities, or both, may be leading to new global climate regimes with widespread effects on food production, energy consumption, and water resources.”487 They recommended that the President appoint an agency—be it NOAA, the NSF, NASA, or whomever—to take the lead in organizing government-sponsored research pertaining to climatic change and the problems it might induce.488

Much of the Congressional interest in the National Climate Program Act came not from believers of the “prophets of doom,” but instead from representatives of largely rural states with concerns over how scientific agencies could meet their practical

484 U.S. Congress, House Committee on Science and Technology, Hearings, The National Climate Program Act, 94th Congress, 2nd Session, May 18-20, 25-27, 1976 (Washington, D.C.: U.S. Government Printing Office, 1976). 485 U.S. Congress, Senate Committee on Commerce, Science, and Transportation, Hearings, National Climate Program Act, 95th Congress, 1st Session, June 8-10 and July 5, 6, and 8, 1977 (Washington, D.C.: U.S. Government Printing Office, 1977), 166 486 Ibid. 487 S. 1652, May 18, 1977, in U.S. Congress, National Climate Program Act, June and July, 1977, 15. 488 Ibid., 14-33.

211 everyday needs as scientific consumers. Members of Congress like Brown and Harrison

Schmitt of New Mexico expressed grave concerns over deforestation, desertification, ice cap depletion, and other large-scale environmental problems associated with climatic change, but most politicians’ primary focus was on the more mundane aspects of climate that affected their states’ farmers and ranchers. They wanted to know about extremes of heat and cold, about storms and severe weather events, and, more than anything else, about water. Climate scientists and Congressional leaders alike saw an immediate need for creating and distributing climatic data to “users” in short- and medium-term agricultural and hydrological planning. “Although human health may be influenced directly by climate,” the authors of Food & Climate argued,

“its greatest impact on man is indirect through its effect on crop

production…Nationwide information and education programs should be

structured to provide farmers with weather and climate information that reduces

the adverse impact of fluctuating weather and climate on farm operations.”489

Ed Epstein of NOAA’s Environmental Monitoring and Prediction unit—and later the head of the National Climate Office—noted that requests for climatic data from the

National Climate Center had climbed from 2,500 per month in 1972 to 5,700 per month during the first quarter of 1977.490

For many scientists, there was more at stake in climate change than the mundane issues of climatic data dissemination for short-term domestic agricultural planning. In developing nations, anomalies of weather and climate could make the difference between survival and starvation. Scientific professionals, they argued, needed to play a stronger

489 National Academy of Sciences, Climate & Food, 3, 11. 490 U.S. Congress, National Climate Program Act, June and July, 1977, 107.

212 role in managing resources like water, food, and energy at a global level. Not surprisingly, the first step in their strategy for influencing public policy was to promote a broad program of coordinated international scientific research. This research would not only address domestic problems of forecasting and climatic variability, however; it would also target the international resource challenges presented by global climatic change.

In the late 1970s, the American Association for the Advancement of Science decided to make climate change the focus of a long-term international initiative in resource management and environmental protection. The largest general scientific society in the world and the publisher of the journal Science, the AAAS was a member- driven advocacy organization committed to advancing science in the public interest.

Unlike the National Academy of Sciences, which responded mostly to the scientific needs of the Congress and the federal government, the AAAS tried to respond to the scientific interests of a larger American public.491 In the late 1960s and early 1970s, the

AAAS had taken a turn to the left, passing a series of resolutions opposing the Vietnam

War and supporting environmental protection, civil and human rights, and gender equality in the sciences. For the most part, however, the organization focused on less volatile practical programs for enhancing international scientific cooperation, improving domestic science education, and increasing the public’s overall engagement with science and technology. In the mid-1970s, the AAAS set up a “Committee on Future Directions”

491 In “The Organization of Scientific Men,” The Scientific Monthly, vol. 14, no. 6, (June 1922), pp. 568- 578, J. McKeen Cattell characterizes the difference between the AAAS and the NAS as a difference between the scientific House of Commons and the scientific House of Lords. In the mid-1970s, this analogy largely held true (as it does today). When AAAS Executive Officer William Carey described the AAAS as a “populist Association” at a meeting in 1978, however, perhaps he overstated the case. “Meeting on Possible AAAS Role in Climate,” March 15, 1978, AAAS Climate Program Records, AAAS Archives, Washington, D.C.

213 in order to refocus these scientific, political, and educational activities. 492 After relatively successful but admittedly ad hoc AAAS participation in United Nations meetings on population, human settlements, and desertification following the U.N. Conference on the

Human Environment, the committee agreed to make CO2-induced climate change a programmatic focus of the organization. Noting that the multi-disciplinary subject of climate represented the sort of “long-range scientific effort of international consequence” that could “command the interest of a wide segment of the association’s membership,” the AAAS established a Committee on Climate as its premier international initiative, with former AAAS president Roger Revelle as its chair.493

Revelle and his colleagues at the AAAS envisioned the organization as a nexus for American scientific, political, and practical interest in climate and climatic change.494

In 1978, the group identified ten federal agencies involved in climatic research: the

Departments of Agriculture, Commerce, Defense, Energy, and Transportation, the

Interior, and State; the Environmental Protection Agency; NASA; and the National

Science Foundation. Under the auspices of the National Climate Program, these agencies had begun to coordinate their research with each other and with state-level agencies and scientific institutions, but they had yet to develop a mechanism for dealing with the various international groups—mostly U.N. agencies—interested in climate change.

These groups included the World Meteorological Organization, the World Health

Organization, the U.N. Environment Programme, and the Food and Agriculture

492 The Committee on Future Directions included, among others, former AAAS President Margaret Mead and planetary astronomer Carl Sagan. “Progress Report on the Committee on Future Directions,” Feb. 8, 1978, AAAS Board Materials for the Committee on Climate, AAAS Climate Program Records, AAAS Archives, Washington, D.C. 493 Revelle served as the organization’s president in 1974. “AAAS Committee on Climate,” undated (1978), AAAS Board Materials. 494 Draft notes of AAAS Advisory Group on Climate meeting, May 26, 1978, AAAS Working Group on Climate Meeting (Transcript), AAAS Climate Program Records.

214 Organization. The AAAS saw gaps between earlier, mission-oriented government projects like CIAP, the limited and often overly academic discussion of the National

Academy of Sciences, and the complex bureaucracy of the on-going international research effort at organizations like UNEP and the WMO.495 Revelle and his advisory group—including prominent members of the climate science community like former

NOAA chief and NAS Climate Research Board member Robert White, NCAR’s William

Kellogg, and Eugene Bierly of the NSF’s Climate Dynamics Research Section, as well as

Congressman Brown and his staff director and science advisor Tom Moss—proposed that the AAAS serve as a link between these various organizations. They hoped the AAAS could build, in White’s words, “the bridge between the complexities of the science and the general public and the policy-makers.”496

The AAAS group emphasized the importance of international cooperation and interdisciplinary communication. “It would be hard to find a problem in science that has such global characteristics,” Robert White contended.497 Data collection and analysis required input and cooperation from scientists studying disparate subjects across the globe, and subsequent discussions about the policy implications of climatic research would necessarily include participation from a diverse community of nations, including many from the Third World.

Revelle’s advisory group went out of its way not just to frame climate in terms of resources, but to define climate science itself as a potential resource for less developed

495 There are actually two drafts of the same meeting in the AAAS Archives, one including the names of speakers and myriad marginal notes, and one, very similar, put in paragraph form with no names or marginalia. I have used the documents in tandem and almost interchangeably, and cite them together as Draft notes of AAAS Advisory Group on Climate meeting, May 26, 1978, AAAS Working Group on Climate Meeting (Transcript), AAAS Climate Program Records, AAAS Archives, Washington, D.C. 496 Ibid. 497 Ibid.

215 countries.498 The AAAS saw climate science as an environmental management tool, much like the sciences of forestry or agronomy. “Climatology is a peculiarly applied science in that its raison d’être is its relation to human welfare,” the group noted, “you have to find out those things about climate that mean something to somebody.”499

Disseminating climatic information, they believed, would provide developing countries the opportunity to better exploit favorable conditions when they existed, and to protect themselves against anomalies when they occurred. Managing climate and climatic information required participation not just from governments and organizations, but also from individuals responsible for managing and utilizing resources at the local level. With a U.N. sponsored World Climate Conference on the docket for 1979, the AAAS hoped to establish better channels of communication between the World Meteorological

Organization and the farmers and water resource managers most affected by the potential impacts of climatic variability and change.500

White and Revelle saw that political solutions to the global problems associated with climate change would require some level of developing-world buy-in, and this, in turn, would require an unusual degree of scientific certainty. Without discussing specific mitigation policies, they noted that preventing climatic change or mitigating its impacts

498 Robert White suggested “Climate as a resource for LDCs” as a potential theme for a proposed NGO event to planned to coincide with the World Climate Conference. In one of his many administrative hats, White served as the Chair of the Organizing Committee for the World Climate Conference—a conference, ironically, that the AAAS Committee on Climate criticized as too narrowly focused on governments and international agencies. It would be difficult to take his suggestion as a strategy to undercut the conference; rather, he, like many scientists, saw the U.N. Conference on the Human Environment as relatively successful, and the concurrent symposia as ultimately beneficial to the larger goals of the U.N. Notes of a Meeting on Possible AAAS Role in Climate, March 15, 1978, AAAS Board Materials; Draft notes of AAAS Advisory Group, May 26, 1978; AAAS Working Group on Climate Meeting (Transcript), AAAS Climate Program Records. 499 As cited in the Progress Report of the Committee on Future Directions, the AAAS mission included a vague mandate to “improve the effectiveness of science in the promotion of human welfare.” Progress Report on the Committee on Future Directions, Feb. 8, 1978, AAAS Board Materials; Draft notes of AAAS Advisory Group, May 26, 1978. 500 Draft notes of AAAS Advisory Group, May 26, 1978.

216 might include potentially disruptive restrictions on behavior and would require heavy capital investment. Committee members saw a need for “an unusually high degree of unanimity and clarity on the part of the scientific community to obtain from political leaders the decisions and actions required.”501 And that unanimity needed to come not only from the atmospheric and geophysical scientists studying the climate itself, but from the social and biological scientists addressing the “higher order effects” of how climatic variability might impact biological, social, and economic systems at various levels.502

The scientific community, argued Tom Moss, should strive for international cooperation in order to create a global scientific consensus on climate and climatic change, because

“unless there is a world science policy base for that kind of change, nothing is going to happen.”503

Moss and others in the Climate Committee advisory group believed that a strong scientific consensus, broadly publicized, would lead to positive political action on climate change. “Political problems,” Moss argued,

“will be moved by the ‘forcing function’ of knowledge. A worldwide scientific

consensus of the possible impact of climatic trends will be the forcing function for

the difficult political decisions, such as regulating land use.”504

Through the “forcing function of knowledge,” scientific consensus, held up in the right governmental and institutional channels, would become a political consensus. By helping foster better international cooperation and coordination of climate science, then, the AAAS would also eventually help make better climate policy.

501 Ibid. 502 Ibid. 503 Ibid. 504 Emphasis in original. Ibid.

217 The AAAS advisory group was sanguine about the efficacy of scientific consensus in part because they saw a consensus already beginning to form on the basic

climatic effects of CO2 within the American climate science community. Reports of the

NAS Climate Research Board (CRB) continued to link CO2 to warming throughout the mid-1970s, and by 1976, only a few scientists still held to the idea of global cooling. As one member of the CRB put it in 1979, “a plethora of studies from diverse sources indicates a consensus that climate changes will result from man’s combustion of fossil fuels and changes in land use.”505 Stephen Schneider had criticized the slow and plodding process of consensus building as a poor way to do science in The Genesis

Strategy, but Moss and his AAAS colleagues now saw an opportunity to capitalize on the power of a unified scientific voice. They hoped to push the process of consensus- building forward by bringing scientists from different disciplines together to hash out those presumably minor points where they disagreed.

In retrospect, the AAAS group’s faith in the power and possibility of consensus seems at best misplaced and at worst hopelessly naïve. As Naomi Oreskes, Erik Conway, and Matthew Shindell explain in their work on Scripps Institute of Oceanography

Director William Nierenberg and what they call the “social deconstruction of scientific

505 As Naomi Oreskes, Erik Conway, and Matthew Shindell point out, the NAS was not the only body confirming a consensus on CO2-induced warming. The “Jasons,” which Oreskes et al. describe as “a secretive group of scientists, mostly physicists, with high level security clearances who have advised the U.S. government on science and technology since the early 1960s,” released a report in April of 1979 concluding that a doubling of atmospheric CO2—the somewhat arbitrary standard taken up by predictive modelers since as early as Svante Arhenius’s pencil and paper equations in the 1890s—would lead to a 2-

3ºC rise in global mean temperatures. CO2 itself, they predicted, would double around 2035. Oreskes et al, “From Chicken Little to Dr. Pangloss: William Nierenberg, Global Warming, and the Social Deconstruction of Scientific Knowledge,” Historical Studies in the Natural Sciences, vol. 38, no. 1 (Winter 2008): 109-152; Jule Charney et al., Carbon Dioxide and Climate: A Scientific Assessment, National Research Council, Ad Hoc Study Group on Carbon Dioxide and Climate (Washington, D.C.: National Academy Press, 1979). This quotation is cited in Oreskes et al. as “Memo, Climate Research Board, Assembly on Mathematical and Physical Sciences, National Academy of Sciences, ‘An Evaluation of the

Evidence for CO2-Induced Climate Change,’ NAS AMPS, Film Label: CO2 and Climate Change: Ad Hoc: General.”

218 knowledge,” scientific consensus, for all its power, can become extremely vulnerable in the face of dissent.506 Scientists rarely define a rubric for establishing consensus on an issue, but frequently, as was the case with the AAAS, unanimity stands as the scientific ideal. As long as an individual scientist can demonstrate her adherence to the standards of “good science”—political neutrality, objectivity, and a disciplinarily defined methodological rigor—her single dissenting voice can undermine the ideal of unanimity and cast doubt upon a scientific community’s “consensus” view. The more powerful or prestigious the dissenting scientist (Nierenberg, as Oreskes et al point out, was quite influential), the greater the challenge to consensus. As both environmentalists and anti- tobacco advocates learned during the 1970s and ‘80s, destroying consensus by manufacturing doubt was far easier than forging even an overwhelming majority of agreement, let alone establishing a unanimous viewpoint.507 And in fact, much of the debate about global warming in the years since 1979 has been a contest over meaning and value of the idea of scientific consensus.

For climate scientists, however, these were mostly lessons of the 1980s. In 1978, their bitterest battles over consensus still lie ahead. With a National Climate Program that found support from both parties in Congress, a Democratic President supportive of renewable energy, and an increasingly interested world public, the AAAS saw an

506 See Oreskes et al., “From Chicken Little”; Oreskes and Conway, “The Denial of Global Warming,” Chapter 6 in Merchants of Doubt: How a Handful of Scientists Obscured the Truth on Issues from Tobacco Smoke to Global Warming (New York: Bloomsbury Press, 2010): 169-215. 507 See Sam Hays, “The Politics of Science,” Chapter 10 in Beauty, Health, and Permanence, 329-362. See also Stephen Bocking, Natures Experts: Science, Politics, and the Environment (New Brunswick: Rutgers University Press, 2006). For more on the role of science and scientific consensus in the tobacco issue, see Robert Proctor, Cancer Wars: How Politics Shapes What We Know and Don’t Know About Cancer (New York: Basic Books, 1995). See also Allan Brandt, The Cigarette Century: The Rise, Fall, and Deadly Persistance of the Produce that Defined America (Boulder, CO: Perseus Books Group, 2007); Stanton A. Glantz, John Slade, Lisa A. Bero, Peter Nanauer and Deborah E. Barnes, The Cigarette Papers (Berkeley: University of California Press, 1996).

219 opportunity to turn good climate science into good climate policy by bringing together the diverse group of biological, physical, environmental, and social scientists working on problems of climate change.

The AAAS found an ally in President Carter’s new Department of Energy. An exceptionally cold winter and a shortage of natural gas in 1976-77, along with ever- increasing gas prices and continued inflation, confronted Carter at his inauguration, and his single term would be marred by energy-related crises from TVA dams and the endangered Snail Darter to the overthrow of the Shah of Iran and subsequent protracted

American hostage situation in that country.508 Though his policies largely failed to protect the American economy from oil price shocks in the short term, Carter’s newly formed Department of Energy began to tackle a wide variety of lasting and important environmental issues in its energy research.509

In 1977, the DoE’s Office of Health and Environmental Research began a collaborative project with the AAAS to study the climatic effects of an increase in

atmospheric CO2. The DoE-AAAS study on CO2—actually a series of interdisciplinary conferences beginning with a workshop in Annapolis, Maryland in the spring of

1979—provided the AAAS Committee on Climate with a working national-level

508 For a concise and informative analysis of Carter’s battle to confront the nation’s energy problems, see John C. Barrow, “An Age of Limits: Jimmy Carter and the Quest for a National Energy Policy,” in The Carter Presidency: Policy Choices in the Post-New Deal Era, edited by Gary M. Fink and Hugh Davis Graham (Lawrence: University Press of Kansas, 1998), 158-223; James T. Patterson, Restless Giant: The United States from Watergate to Bush v Gore (New York: Oxford University Press, 2005), 125-6. 509 Carter’s energy policies clearly failed within the tenure of his presidency, and both Barrow and Patterson at least partially pin his 1980 electoral defeat on his inability to control energy prices in an extremely volatile energy economy. Barrow does suggest, however, that the energy glut that followed during Reagan’s eight years in office had at least a little to do with Carter’s controversial policies in the late 1970s, and that in the sense that they helped stabilize energy prices for American consumers (though at a much subsidized rate), Carter’s energy program actually worked. Barrow, “An Age of Limits,” 176. As Sheldon Ungar demonstrates in “The Rise and (Relative) Decline of Global Warming as a Social Problem,” former Nixon CEQ member Gordon MacDonald and Alvin Weinberg, who coined the term “big science,” had a role in creating a CO2 research program at the Department of Energy. In Sociological Quarterly, 33 (1992), pg. 488. Cited in Oreskes et al., “From Chicken Little to Dr. Pangloss,” 115.

220 example of the kind of conference they hoped to sponsor at an international level. It had ample funding, widespread participation from the scientific community, and the promise of a direct to policy-making within an important executive agency.

With 85 scientists from a wide variety of fields in attendance, the workshop was the first step in creating just the sort of scientific clearing-house the AAAS Committee on

Climate envisioned. Organized chiefly by the DoE’s self-described “high priest of CO2,”

David Slade, and AAAS Climate Project Coordinator David Burns, the Annapolis workshop broke climate research down by subject into five major panels.510 They were:

“Environmental Effects of the Oceans, Cryosphere, and Ocean Biota,” “Environmental

Effects on the Less-Managed Biosphere,” “Environmental Effects on the Managed

Biosphere,” “Social and Institutional Responses,” and “Issues Associated with Analysis of Economic and Geopolitical Consequences.”511 In contrast to the National Academy of

Sciences studies of 1976 and 1977, the DoE-AAAS workshop didn’t shy away from the

political implications of CO2-induced climate change. To an extent, Revelle and “the

Davids” embraced them, even inviting Mickey Glantz to deliver a paper entitled “A

512 Political View of CO2.” The workshop’s participants identified most of the major issues that would continue to dominate discussions about global warming and climatic change for the next three decades.

Environmentalists, however, were once again conspicuously absent from the climate discussion. Despite the enthusiasm of the scientific community, American

510 Slade described himself as the “high priest of CO2” in jest in David Slade to David Burns, October 2, 1980, AAAS Climate Program Records. 511 Executive summary, AAAS/DOE Workshop on “Environmental and Societal Consequences of Possible

CO2-induced Climate Change,” Annapolis, Maryland, April 2-6, 1979, AAAS Climate Program Records, 512 Michael Glantz, “A Political View of CO2,” a discussion paper presented at the DOE-AAAS Workshop on Environmental and Societal Consequences of Possible CO2-induced Climate Change, Annapolis, Maryland, April, 1979, AAAS Climate Program Records.

221 environmental organizations showed little interest in participating in the joint AAAS-

DoE workshop on CO2. To many environmentalists, the DoE, an organization that managed the nation’s nuclear energy and had at least partial jurisdiction over many of its dams and coal-fired power plants, was the enemy.513 Proposing to invite representatives from the Natural Resources Defense Council, FoE, the Sierra Club, and the

Environmental Defense Fund, Slade noted to Burns in November of 1978 that “one or two people should come to represent the ‘environmentalists.’ Groups such as above have

514 appeared totally disinterested in the CO2 question.” With the exception of Thomas

Kimball of the National Wildlife Federation, who lodged his official support for the

National Climate Program Act in a letter to Congress, environment groups generally assigned the national effort to study the problem of climate change a low priority and remained somewhat aloof from the climate debate into the early 1980s.515

Conclusion

Environmentalists’ tepid initial response to the issue of climate change underscored a larger set of philosophical differences between professional environmentalists and politically active scientists over both the tactics and priorities of environmental protection. In the 1970s, American environmentalists thought primarily in terms of “protection,” and their institutions reflected their defensive mindset. Beginning in the early 1970s with Sierra Club Legal Defense Fund and the Natural Resources

Defense Council, environmental organizations began to focus the lion’s share of their

513 Before the DOE began collaborating with the AAAS on the Annapolis conference, they had held a conference of their own at Oak Ridge National Laboratories—a highly controversial nuclear research facility. Oreskes and Conway, Merchants of Doubt, 171. 514 David Slade to David Burns, November 11, 1978, AAAS Climate Program Records. 515 U.S. Congress, National Climate Program Act, June and July, 1977, 466.

222 energies on Congressional lobbying and legal action in response to immediate affronts to local and regional environments.516 This was an essentially reactionary strategy that in retrospect stands in stark contrast to the preventative approach of the scientists working with the Department of Energy and the AAAS. Laboring diligently to influence new legislation and uphold the standards of existing legislation in order to counteract existing corporate and governmental environmental transgressions, environmentalists planned future strategies of environmental protection based on existing or immediate environmental hazards. Because of their uncertainties, what McCloskey referred to as

“first-generation issues” like increased atmospheric CO2 and ozone depletion that had yet to germinate into tangible threats could support little in the way of strategic planning or fund-raising.517

Concerned climate scientists, on the other hand, worked within rather than against executive bureaucracies. They were generally unburdened by the immediate needs of public constituents, and they sought to illuminate the potential future risks of what Glantz referred to as “those low-grade, but continually increasing, insults to the environment for which pluralistic society (the national as well as international community) has not yet found an effective policy-making process.”518 Rather than reacting to the impacts of climatic change, many of these scientists believed, governments and societies should work to preempt environmental catastrophe by mitigating and adapting to potentially disastrous changes still in the future. Theirs was not the “gospel of efficiency,” but the

516 NRDC, coincidentally, was among the first national-level environmental organizations to incorporate climate change into its programs for the 1980s. See Chapters 5 and 6. 517 “Draft: Sierra Club International Committee Meeting of June 13-14, New York/Terrytown (1980),”Meetings and Conferences, International Committee, 1972-1983, Sierra Club International Program Records [3:18]. 518 Michael Glantz, “A Political View of CO2.”

223 gospel of preparedness.519 Through the National Climate Program Act, the National

Academy of Sciences, and the joint AAAS-DoE workshops, climate scientists worked to incorporate climate change into government policy from the inside out.

Scientists and environmentalists also differed over the relative priorities of economic development and environmental protection. Echoing the early 20th Century debate between “conservationists” like Theodore Roosevelt and his Forest Service Chief

Gifford Pinchot and “preservationists” like Sierra Club icon John Muir, environmental leaders of the 1970s like McCloskey, David Brower, and Thomas Kimball saw a fundamental tension between rapid economic development and the health of the natural environment.520 “While I think we should be very sensitive and understanding with respect to the need for economic advancement by the poverty-stricken,” wrote

McCloskey shortly after the 1972 U.N. Conference on the Human Environment,

“I think that we are not in a position to be emphatic in stating that the two goals

are entirely compatible. On the contrary, I think we see some very clear problems

with respect to reconciliation, though I think we can express every hope that we

can find a way to have both a tolerable environment and sufficient human

economic security.”521

Many of the scientists most vocal in debates about the impacts of climate change, meanwhile, shared a neo-Progressive commitment to socially and environmentally responsible resource management that, in their eyes, could reconcile these tensions in the

519 Samuel P. Hays, Conservation and the Gospel of Efficiency: The Progressive Conservation Movement, 1890-1920 (Cambridge: Harvard University Press, 1959). 520 Ibid. 521 Michael McCloskey to Patricia Rambach, August 15, 1972, Stockholm Conference on the Human Environment, 1972-73, Subject Files, 1972-1985, Sierra Club International Program Records [5:20].

224 service of an overall improvement in the quality of human life.522 Despite the dire warnings embedded in their conclusions, studies by Roger Revelle, Mickey Glantz,

Stephen Schneider and others rested on the more optimistic premise that with good scientific research behind them, policy-makers could—and should—encourage new types of development in order to protect the coupled human and natural systems potentially affected by irresponsible—or “unscientific”—growth.

During the 1970s, climate scientists gained remarkable access to federal agencies and policy elites, and they began to build the type of consensus Revelle and his AAAS colleagues had hoped for. In 1979, the National Academy of Sciences committee, headed by meteorologist Jule Charnay of MIT, released a comprehensive assessment of work

done by scientists over the previous decade on CO2 and climate. The study concluded that a doubling of atmospheric CO2 would lead to a 1.5-4°C warming of the Earth, and

that if atmospheric CO2 continued to increase, there was “no reason to believe that these changes will be negligible.”523 The joint AAAS-DoE venture convinced David Slade that the Department of Energy should not only continue to fund further research on the effects

of CO2-induce climate change, but should begin to develop a national plan for ameliorating and adapting to the unintended climatic consequences of fossil fuel energy consumption, beginning as early as 1979.524 In a detailed series of flow charts, he laid out a potential research and assessment program that would support a continuing dialogue between climate scientists and policymakers through the end of Carter’s theoretical

522 Hays, Conservation and the Gospel of Efficiency. 523 Cited in Orskes et al., “From Chicken Little to Dr. Pangloss,” but comes from Verner Suomi’s introduction to Jule G. Charnay et al., Carbon Dioxide and Climate: A Scientific Assessment, National Research Council, Ad Hoc Study Group on Carbon Dioxide and Climate (Washington, D.C.: National Academy Press, 1979), vii. 524 “Action Flow, U.S. Carbon Dioxide Research and Assessment Program,” from David H. Slade (no date), Workshop Correspondences, 1979 AAAS/DOE Workshop File, AAAS Archives, Washington, D.C.

225 second term. In 1979, Nature called climatic change “the most important environmental issue in the world today,” and Slade sought to establish a prominent place for the issue in

America’s ongoing discussion about its most pressing political and economic problem: energy.525 Working within government, climate scientists appeared to be on their way to

creating meaningful policies on CO2 and climate change for the 1980s.

But climate scientists’ commitment to working within government bureaucracies left them vulnerable to political change. Slade’s program, like Carter’s second term, was not to be. Despite strong talk of a responsible renewable energy program, Carter’s

troubles in the Middle East led him to back CO2-producing synthetic fuels made primarily from coal. Carter’s commitment to synthetic fuels put Slade in an awkward position vis-à-vis Presidential energy policy. When Ronald Reagan was inaugurated,

Slade’s influence within the DoE decreased even further. Amidst severe budget cuts to government-funded social science and environmental research instituted by the new

President, the director of the DoE’s Division of Carbon Dioxide and Climate Research found himself isolated within an agency whose new mandate under Reagan would be

“cheap energy now!.” “I am not paranoid,” Slade quipped to David Burns in the fall of

1980,

“but, besides wrapping my head in silver foil before retiring so as to ward off the

rays the CIA beams at me nightly, I carefully review publications sent to me to

see how ‘they’ are cutting off my information sources….As the high priest of CO2

for the federal government I view it as more than just a coincidence that my copy

526 of science [sic] is missing the article on CO2.”

525 “Costs and Benefits of Carbon Dioxide,” Nature, Vol. 279 (May 3, 1979): 1. 526 David Slade to David Burns, October 2, 1980, AAAS Climate Program Records.

226 When Reagan took office, the writing was on the wall; within a year, Frederick

Koomanoff, a loyal Reaganite who earned the nearly universal disdain of global warming activists in the years to come, took Slade’s place, completely reshaping the DoE’s

position on CO2 and climate change. In the early 1980s, concerned climate scientists suddenly found their access to government elites curtailed, and many had their federal research funding slashed.

Recalling their experiences three decades later, few of the scientists involved in the AAAS-DoE conferences and the National Climate Program remember their involvement in scientific discussions of climate change as environmental or political activism. Schneider’s The Genesis Strategy notwithstanding, climate scientists made few appeals to the broader public, and the research programs they advocated did not yet have significant non-scientific political components. But behind climate scientists’ push for more “good science” rested a precautionary ethos consistent with American environmentalists’ persistent concerns over the unintended consequences of human

actions. By their own formulation, climate scientists’ efforts to build consensus on CO2- induced warming and their advocacy for further research implied support for the yet undefined regulations and government programs their science would, through the

“forcing function of knowledge,” underpin. Their “gospel of preparedness” was at once scientific and political. As the 1980s would show, so too was the controversy it engendered.

227 Chapter 5 Scientists, Environmentalists, Democrats: The New Politics of Climate Change Under Ronald Reagan

America’s professional environmentalists took on the issue of global warming slowly and haltingly in the early 1980s, but by 1985 most of the “group of ten” had at the very least begun to recognize climate change as a potential issue in their organizations’ list of national and international priorities for the future. In part, the confluence of climate science and American environmentalism arose out of common interests and common concerns established in the 1970s. During that decade, national-level environmental groups like the Sierra Club and Friends of the Earth began to develop more robust strategies for international environmental protection, and—Michael

McClosky’s narrow guidelines for environmental campaigns notwithstanding—they increasingly began to focus on environmental degradation at a global scale.527 Climate scientists, meanwhile, increasingly began to focus on the potential social and

environmental impacts of CO2-induced warming, including impacts on species, habitat, and natural resources. In addition, as scientists continued to investigate the complexities

of CO2, they began to address the climatic impacts of phenomena in the biosphere, and particularly of deforestation—a hot topic for environmentalists.528 Perhaps most importantly, both communities harbored deep concerns about the wide-ranging impacts

of non-renewable energy use from oil, coal, and gas, and both groups saw CO2 as one in a

527 Michael McCloskey, “Criteria for International Campaigns,” December 1, 1982. International Committee, 1972-1983, Meetings and Conferences, Operational Files, Sierra Club International Program Records [3:19]. See chapter 4. 528 See G.M. Woodwell, R.H. Whittaker, W.A. Reiners, G.E. Likens, C.C. Delwiche, and D.B. Botkin, “The Biota and the World Carbon Budget,” in Science, Vol. 199 (January 13, 1978), 141-146.; George Woodwell, “The Carbon Dioxide Question,” Scientific American, Vol. 238, No. 1 (January, 1978): 34-43. See also Weart, 103-107.

228 long list of energy-related environmental problems that they believed should help to guide federal energy policy in the 1980s.

Common scientific and environmental interests continued to drive these groups’ relationships in the early 1980s, but in the new decade climate scientists and American environmentalists also came together to face a common political enemy. His name was

Ronald Reagan. Elected in 1980, the new President harbored a disdain for what he saw as an unnecessarily alarmist and anti-business environmental movement. He coupled this disdain with a similar contempt for social and environmental science research. Upon inauguration, he sought to dismantle both the CEQ and the DoE, and he replaced capable administrators at the EPA and the Department of the Interior with zealous political acolytes hostile to the idea of environmental regulation.529 The Administration aggressively downsized federal environmental research, focusing in particular on research into renewable energy. Unable to fully excise the DoE from the executive branch, Reagan’s Office of Management and Budget reallocated more than a billion dollars of the DoE’s money to the Department of Defense.530 The ham-fisted approach

531 left the DoE with neither the funding nor the personnel to continue its research on CO2.

Like environmental scientists of every stripe, scientists involved in the joint AAAS-DoE

529 Michael Schaller, Reckoning with Reagan: America and Its President in the 1980s (Oxford: Oxford University Press, 1980), 50; See also “The Reagan Antienvironmental Revolution” in Samuel P. Hays, Beauty, Health, and Permanence: Environmental Politics in the United States, 1955-1985 (Cambridge: Cambridge University Press, 1987), 491-526. 530 Stephen H. Schneider, Science as a Contact Sport: Inside the Battle to Save Earth’s Climate (Washington, D.C.: National Geographic, 2009), 90. See also U.S. Congress, House Committee on Science and Technology, Hearing, Fiscal Year 1983 Department of Energy Budget Review (Fossil Energy and Basic Research), 97th Congress, 2nd Session, March 18, 23, 24, 25, 30, 1982 (Washington, D.C.: U.S. Government Printing Office, 1982), 1. 531 Schneider, Contact Sport, 90. In a 1982 hearing, an exasperated Congressman James Scheuer lamented that “They have cut us down to $4 million, approximately the cost of two M-1 tanks that don’t work very well, approximately one quarter of one day’s military budget.” U.S. Congress, House Committee on Science and Technology, Hearing, Carbon Dioxide and Climate: The Greenhouse Effect, 97th Congress, 2nd Session, March 25, 1982 (Washington, D.C.: U.S. Government Printing Office, 1982), 7.

229 CO2 project begun under Carter —the largest and most extensive program of climate research up to that point—saw their funding slashed and their reliable government contacts replaced by party-line political appointees. Having worked hard to incorporate climate change into the research and policy agendas of federal agencies since the 1950s, climate scientists suddenly and unexpectedly found themselves on the outs. As a result, in the 1980s, the politics of global warming became, like environmental politics more broadly, a politics of dissent.

For climate scientists, the transition from working within government to working against it was abrupt and traumatic, and their collective dissent took many forms. Some

government climate scientists used CO2-induced warming to attack the Administration’s policies on energy and energy-related research from within the bureaucracy, with mixed results. They met stiff resistance from Administration officials who sought to use their control over the mechanisms of the executive bureaucracy to manage the content of government-sponsored scientists’ conclusions. These and other outspoken scientists like

Stephen Schneider and Roger Revelle also began to align with Congressional Democrats who objected to the President’s energy policies, and who hoped to pillory him for trying to dismantle the Department of Energy. Scientists used Democrats’ hearings on energy and climate as a way to push back against the Administration’s cuts in climate research

and to raise the public profile of the issue of CO2-induced climate change. Because of the focus on energy and fossil fuels, the hearings also attracted leaders from environmental

groups interested both in promoting future CO2 research and in criticizing the

Administration’s energy and environmental policies more broadly. Scientists, environmentalists, and some congressional Democrats had already established loose

230 political alliances based on common interest in the late 1970s; under Reagan, these loose affiliations grew into a form of coordinated political opposition.

Reagan’s opponents initially used climate research to attack the President’s energy and energy research policies, but left-leaning climate scientists soon mobilized their expertise in response to his aggressive foreign policy as well. They focused particularly on the Administration’s national defense strategy. Aghast at Defense

Department officials’ talk of a “winnable” nuclear war and unimpressed with Reagan’s

Strategic Defense Initiative (SDI), these scientists used climate models to predict the potential climatic impacts of a nuclear exchange.532 Their results, which foretold of a dramatic decrease in global temperatures caused by ash, smoke, and dust—a “”—fit into environmentalists’ ongoing efforts to illuminate the broader environmental impacts of nuclear weapons. Led by popular astrophysicist Carl Sagan and biologist Paul Ehrlich, liberal-minded scientists argued that the potential severity of a nuclear winter provided new impetus for aggressive international arms control policies in line with those of the Nuclear Freeze movement.533 Their work further reinforced the

532 Schneider, Science as a Contact Sport, 98; Richard Halloran, “Pentagon Draws Up First Strategy for Fighting a Long Nuclear War,” The New York Times, May 30, 1982; Richard Halloran, “Weinberger Confirms New Strategy on Atom War,” The New York Times, June 4, 1982; Tom Wicker, “Crossing a Thin Line,” The New York Times, October 21, 1981. See also Caspar Weinberger, Fighting for Peace: Seven Critical Years in the Pentagon (New York: Warner Books, 1990). 533 MAD is a doctrine of military strategy derived from the doctrine of “deterrence” that has stood at the heart of American defense policy since advent of the atomic bomb. Adherents to the theory of deterrence hold that the existence of weapons of sufficient threat to an enemy can, through the fear generated by the magnitude of the consequences of using those weapons, ultimately prevent their use. MAD is a specific type of deterrence in which the full-scale use of nuclear weapons by two opposing sides would result in the destruction of both the attacker and the defender. See John Lewis Gaddis, Strategies of Containment: A Critical Appraisal of Postwar American National Security (Oxford: Oxford University Press, 1982). In Essence of a Decision: Explaining the Cuban Missile Crisis, Graham T. Allison disputes the MAD doctrine. MAD, he argues, is based on a “rational actor” model of foreign policy in which indiviual policymakers make consequence driven decisions based on the best available information. As he shows through his “organizational process model” and his “bureaucratic process model,” however, this is not always the case. For early usage in American foreign policy, see Robert McNamara, “Mutual Deterrence,” a speech by the Secretary of Defense, September 18, 1967, San Francisco, CA. For current applications of

231 climate science community’s relationships with professional environmental groups and certain Congressional Democrats, both of whom greeted Reagan’s nuclear defense policies with skepticism.534 In addition, the issue provided a demand for more climate research in an era of deep cuts in government funding, a boon to the climate science community at large. On the other hand, Sagan’s and Erhlich’s work also led to an embarrassing set of public disagreements within the atmospheric science community over the science of nuclear winter. These disagreements played out not only in scientific journals, but also in the press and in the pages of Foreign Affairs. In the end, concern over nuclear winter provided funding and public visibility for climate research, but the ensuing debate also undercut the credibility of scientific activists and embittered partisans on both sides of a deepening political divide within the atmospheric science community.

Climate scientists’ responses to the “Reagan Reaction” ultimately changed the nature of the climate change issue in the 1980s. Since the 1950s, when individual

scientists worked to incorporate CO2 and climate into the Cold War research agenda, the study of climate had always involved a certain engagement with national politics. Up

through the 1970s, the politics of CO2 revolved primarily around questions of science policy and, more importantly, science funding. In the 1980s, this effort to influence federal science policy persisted. But climate scientists also began to associate their work more closely with specific energy and defense policies in the 1980s. Scientists’ conclusions increasingly bled into policy proscriptions, a challenge to the vaguely-

MAD and the fate of the doctrine since the end of the Cold War, see Keir A. Lieber and Daryl Press, “The Rise of U.S. Nuclear Primacy,” Foreign Affairs, March/April 2006, pg. 42-55. 534 In the United States especially, environmentalists coupled their opposition to nuclear weapons with deep suspicions about nuclear energy—fears reawakened by the 1979 Three Mile Island Disaster. Interestingly, because nuclear power is not a greenhouse polluter and thus served as a likely alternative to fossil fuels,

Three Mile Island also served to complicate environmentalists’ involvement in the CO2 issue as it related to energy in the early 1980s. It continues to haunt the discussion today.

232 defined boundaries between science and policy that many scientists argued were essential to the credibility of the profession. Moreover, in marrying their research to the issues of energy and defense—two of the most hotly contested issues in American politics at the time—politically active climate scientists raised the stakes of what had previously been primarily a science policy debate. As a result, the public politics of climate change became more partisan and more combative; in the process, so too did the science.

The Transition

Few individuals’ stories capture the rapid and traumatic change that accompanied the transition from Reagan to Carter more completely than the experiences of the Solar

Energy Research Institute’s (SERI) Denis Hayes. A jack-of-all-trades within the environmental movement, Hayes had served as one of the key organizers in Senator

Gaylord Nelson’s 1970 Earth Day initiative. Building on the combination of a deep- seated interest in environmental issues and a degree in engineering from Stanford, he soon established himself as one of the nation’s top experts in renewable energy. In 1977

Hayes wrote Rays of Hope: The Transition to a Post-Petroleum World, a book that soon impressed Carter’s Republican Secretary of Energy James Schlesinger. 535 With boosters of nuclear energy at the head of the DoE nuclear program and men passionate about coal

535 Exactly how Schlesinger came to read Rays of Hope is an interesting story in and of itself. During his book tour, Hayes was introduced to the Saudi Oil Minister, Zaki Yamani, one of the most powerful men in Saudi politics. According to Hayes, their discussion was nothing particularly memorable, but out of form he sent the minister a signed copy of the book. Not long afterward, Schlesinger, on a trip to Saudi Arabia, found himself in a waiting room outside Yamani’s office. Bored, and with no western magazines or newspapers to read, Schlesinger picked up the only piece of writing in English available…Hayes’s book. When the DoE began to look for a new director for SERI a few months late after Paul Rappaport had to bow out of the role for personal reasons, Hayes’s names was on the Secretary of Energy’s mind. Denis Hayes, interview with the author, Seattle, Washington, June 6, 2009; Denis Hayes, Rays of Hope: The Transition to a Post-Petroleum World (New York: W.W. Norton and Company, 1977). The copyright to Rays of Hope is actually held by the Worldwatch Institute, an environmental group focused on international problems—particularly the food crisis—then headed by Lester Brown and Eric Eckholm.

233 leading the fossil fuels program, Schlesinger was intent on putting someone passionate about renewables at the head of the DoE renewables program. When SERI needed a new director shortly after opening its doors in 1977, Schlesinger looked to Hayes.536

As one of the institution’s first tasks, Schlesinger’s deputy secretary John Sawhill asked Hayes to put together a study on what the United States could reasonably expect to accomplish with energy conservation and renewable energy production in terms of the nation’s overall energy use. Headed by SERI’s assistant director Henry Kelley, the study concluded that by the year 2000, an aggressive renewable energy policy could account for more than a quarter of the total energy production in the U.S., and that the Carter

Administration should commit to a goal of producing 20% of the total U.S. energy pool—a pool that could be made productively shallower, the report argued, through conservation—through sources like solar and wind power.537 In short, SERI showed the

DoE that renewables were a feasible option for the future. The institute delivered the report, known to its authors alternatively as the Solar Conservation Report or the

Sawhill/SERI report, to the Department of Energy in November of 1980. In the wake of the 1980 election, however, the DoE tabled the study, and its conclusions were at least temporarily forgotten.538

To Hayes, the election at first looked like little more than a bump in the road for renewables. Like many environmentalists, Hayes had moderate hopes for the incoming

Reagan administration—or at least he “wasn’t sure things would be getting a whole lot

536 National Renewable Energy Laboratory, 25 Years of Research Excellence, 1977-2002 (Boulder, CO: NREL, 2002), at www.nrel.gov. 537 Hayes interview, 6/2/09; Solar Energy Research Institute, A New Prosperity: Building a Sustainable Future (Andover, MA: Brickhouse Publishing, 1981). See also Eric Hirst, “Review: A New Prosperity: Building a Sustainable Energy Future by Solar Energy Research Institute,” Policy Sciences, vol. 14, no. 2, Self-Governance in the Interpenetrated Society (April, 1982): 198-202. 538 Hirst, “Review,” 199.

234 worse.”539 Carter had backed off considerably on environmental policy in the waning years of his term, and nowhere more than in his energy policy. In August of 1979, he replaced Schlesinger with Charles Duncan, a Texas millionaire and member of the Coca-

Cola Company’s board of directors who, in Hayes’s view, “had essentially no knowledge about renewables—or anything else related to the environment for that matter.”540 In a series of short radio addresses during the campaign, Reagan specifically lauded renewables as a way to support the libertarian desires of “the great American loner,” and up until Reagan started picking his cabinet, Hayes and others thought they might be able to work with the new President.541 In any case, SERI and the DoE had four years of solid growth propelling them forward; even without a supportive president, Hayes and his colleagues felt secure in their posts.

They were sorely mistaken. The new administration slashed SERI’s budget by nearly 80%, and gave scientists who had left tenured positions at universities to take their

SERI posts two weeks notice.542 The President began replacing officials both in SERI and the DoE with loyal Reaganites. Most egregious was James Edwards, a dentist and former South Carolina Governor who Reagan tapped as the new Secretary of Energy in order to appease new Republicans from the traditionally Democratic South who were upset that the President had no other Southerners in his cabinet.543 For Hayes and his

SERI colleagues, the DoE suddenly looked like a part of the energy problem rather than a potential part of the solution.

539 Hayes interview, 6/2/09. 540 Ibid. 541 Hayes interview, 6/2/09. 542 Hayes interview, 6/2/09. 543 Like many of the scientists and environmentalists that I have interviewed, Hayes still gets visibly upset talking about Jim Edwards’ appointment. “It quickly became clear that Charles Duncan was actually pretty competent,” Hayes remembers, “I mean, Jim Edwards was just fucking unbelievable.” Hayes interview, 6/2/09.

235 Still sitting on the yet unpublished Solar Conservation Report, Hayes and Kelley decided to fight back. Hayes’s memory runs like a scene from a detective novel:

“So I’m back in Washington, D.C. early in 1981 for a meeting with this acting

assistant secretary in the Department of Energy, a guy named Frank DeGeorge. I

always go fishing for intelligence, so I went over to the department an hour earlier

and I was kind of walking around and a friend of mine—one of the deputy

assistant secretaries—stopped me at his desk. And he says, ‘Hey, have you talked

to DeGeorge yet?.’ And I say, ‘no, why?” And he says ‘the Solar Conservation

Report? Yeah, it’s dead as a doornail, they’re never going to let that out.’ So I

strolled down the hall, nipped into the nearest elevator, went back to my hotel,

called Henry Kelley and told him to Xerox as many copies as he possibly could,

but get at least 50 or 100 of these things fed-exed out to environmental

organizations and essentially anyone who might be interested. I had my secretary

call this Frank DeGeorge’s office and say I was sorry I was going to have to

cancel that meeting because I’d come down with the flu, but we’d like to

reschedule it for the next time I’m in town. Didn’t seem to be anything terribly

important; he hadn’t told me it was. So I flew back to Colorado, and low and

behold I get a call from Frank DeGeorge, and we’re talking about the report, and

he says it’s time to pull the plug on it. I said ‘geez, Frank, there’s no way to do

that. I’ve got a hundred of these things out there for review right now.”544

In the short run, Hayes’s and Kelley’s ploy worked. With the report already in the hands of a number of environmental groups and key politicians, Hayes had seen that the

544 Hayes interview, 6/2/09.

236 report made it to its intended audience. Hayes also made sure to leak a copy to

Representative Richard Ottinger of New York, a member of the House Committee on

Energy and Commerce and a strong supporter of federal conservation and solar programs. Ottinger, in turn made it public domain by reading it into the Congressional

Record. Brickhouse Publishing then published the report as A New Prosperity: Building a Sustainable Future in April of 1981.545 Not surprisingly, Hayes was invited to resign for “not being a team player,” but he managed to leave with an effective parting shot.546

His well-orchestrated retaliation aside, Hayes’s fall from government grace typified scientists’ and environmentalists’ experiences throughout the federal government. For climate science, the most important changes occurred in the

Department of Energy, where the appointments radiated outward from Jim Edwards.

Initially, the Reagan Administration explored the option of dismantling DoE altogether, eliminating a wide variety of its programs and rolling what would be left of its core, an entity called the Energy Research and Technology Administration (ERTA, suspiciously similar to ERDA of the 1970s), into either Commerce or Interior.547 Reagan’s aides eventually preserved the DoE as a single, cabinet-level entity, but the Administration cut

545 Ibid.; Hirst, “Review,” 199. 546 Hayes’s dismissal from SERI is also an interesting story. As a college student at Stanford, Hayes had led the occupation of some engineering buildings in 1969, much to the chagrin of then governor Ronald Reagan, who had the National Guard on standby in Redwood City. According to Hayes, when Reagan’s White House Council and former gubernatorial campaign manager Ed Meese associated the name of the wayward SERI director with the brash college student, he flew off the handle. Meese reportedly had a 3X5 card with Hayes’s name and picture on it that he carried around until Hayes was finally fired—or asked to resign—for “not being a team player” in the summer of 1981. Hayes interview, 6/2/09. 547 ERDA was the Energy Research and Development Administration, a product of the Energy Reorganization Act of 1974. Though the various administration’s functions and responsibilities changed slightly over time, ERDA succeeded the Atomic Energy Commission and preceded the Department of Energy. James Watt to the President, “Organization of Energy Functions,” December 10, 1981, George A. Keyworth, III Papers, Department of Energy, 1981-82, Folder 2, Box 3, Ronald Reagan Archives, Ronald Reagan Presidential Library, Simi Valley, California.

237 the DoE’s total budget by more than a billion dollars.548 The Administration’s DoE budget request for FY 1983 represented an 87% reduction from the sum requested by

Carter for FY 1981.549 These budget reductions severely curtailed the research of the

DoE’s Office of Health and Environmental Research, including David Slade’s joint

550 AAAS-DoE CO2 research program. In 1981, the Administration replaced the forward thinking Slade with Frederick Koomanoff, a capable science administrator but a man not

particularly interested in pushing CO2 research in the face of the larger trends in

Administration policy.

Cuts in DoE funding made waves throughout the federal research structure. Most immediately, the Reagan Administration diminished the resources available for Slade’s

AAAS/DoE CO2 project. But the cuts also stressed other sources of government funding for climate research. For example, DoE ran a number of national laboratories—laboratories like Lawrence Livermore, SERI, and Oak Ridge—and researchers at these laboratories scrambled to replace DoE money with funding from the

Department of Defense ($3.8 billion richer in R&D funding under Reagan), the National

Science Foundation, and the private sector. For climate research in particular, this put

548 Somewhat ironically, nuclear energy helped to save the DoE. As Secretary of the Interior James Watt argued, separating the nation’s nuclear energy resources from the whole of Reagan’s energy package left nuclear vulnerable to political attack. Instead, he urged that “nuclear energy should be so interwoven with the total energy package that it cannot be separately attacked and torpedoed.” For Watt, that did not mean keeping the DoE, but the Administration soon realized that giving DoE full oversight of nuclear could accomplish Watt’s goal. Watt to the President, “Organization of Energy Functions,” 12/10/81. Schneider, Contact Sport, 90. 549 U.S. Congress, Fiscal Year 1983, 1. 550 In Science as a Contact Sport, Stephen Schneider claims that during Slade’s presentation of his CO2 program to the incoming DOE officials, Jim Edwards leaned over to his deputy director of research, N. Douglas Pewitt, and said, loudly enough to be heard, “We were put in here to get rid of environmental and social science research, so just forget this project.” The source of the “ear-witness” account remains a mystery, but Pewitt’s actions in public hearings in 1981 and 1982 seem to support the claim; he made every effort to diminish the status of and funding for CO2 and other environmental research within the DoE. Schneider, Contact Sport, 90. For more on the AAAS-DoE program and on Slade’s dismissal, see chapter 4.

238 exceptional strain on the National Science Foundation, which already funded the quasi- governmental NCAR and a number of other university-based climate projects. Worse still for climate scientists, the Reagan Administration ultimately controlled most of these resources, too, and what little new money the NSF could make available for climate research began to go to projects investigating the direct impacts of rising atmospheric

CO2 on crops and managed forests—a set of studies that in isolation made CO2 buildup look a little less threatening, and possibly even beneficial.551

Federal environmental agencies endured similar cuts in funding and changes in personnel. As political appointees, Carter’s Council on Environmental Quality expected to be replaced, and they were. But as it had with DoE, the Reagan Administration first considered disbanding the CEQ altogether, and only relented when Senate Republicans, not wishing to lose a cabinet post, objected.552 In the end, the CEQ, headed by a

551 Schneider, Contact Sport, 87. The connection between USDA and climate change is an interesting one, in part because in an era of overall cuts in CO2 research, in the 1980s research on CO2 and agriculture was allowed to persist. Ultimately, as global warming entered American’s mainstream political consciousness in the late 1980s, this ongoing USDA research gave the outgoing Ronald Reagan and, more importantly, the incoming George H.W. Bush a way to credibly express concern over climate change without directly giving in to calls from Democrats to recognize the problem and do something about it. See U.S. Congress, Senate Committee on Agriculture, Nutrition, and Forestry, Global Climate Change Prevention Act of 1989, Hearing, 101st Congress, 1st Session, November 6, 1989, (Washington, D.C., U.S. Government Printing Office, 1989). Special thanks to Heather West for her undergraduate research on this topic in her unpublished “Agriculture, Climate Change, and Ronald Reagan,” June 8, 2007. 552 Within the Administration, the disbanding of CEQ caused some consternation, in part because the interim head of CEQ, Malcolm Baldwin, hoped to keep the group around, and in part because doing away with the body presented a potentially sticky legal question that the Administration didn’t want to deal with. As Baldwin pointed out, though Nixon created CEQ by executive order, he did so to fulfill a Congressional mandate put forward in the National Environmental Policy Act. CEQ—or some body like it responsible to report to the President on environmental matters but also independent of other federal agencies—was thus statutorily mandated. Reagan’s aides hoped to pass responsibility for NEPA from CEQ to the Department of Interior, which to some extent they did, but because Interior also oversaw energy and natural resources, that move technically violated NEPA itself. The CEQ, though weakened and marginalized, was thus preserved. Baldwin, on the other hand, was soon replaced. Malcolm Baldwin, “Memo for the Vice President and James Baker, III,” February 21, 1981; F. Khekouri to Frank Hodsell, “Budget: Council on Environmental Quality,” March 5, 1981; and Memorandum, Alan Hill to Frank Hodsell, April 16, 1981, all in James A. Baker, III Files, Council on Environmental Quality, Folder 2, Box 1, Ronald Reagan Archives, Ronald Reagan Presidential Library, Simi Valley, California. See also Hays, Beauty, Health, and Permanence, 494.

239 Republican party leader named Alan Hill, saw its staff reduced by half and its funding slashed by two-thirds.553 The deepest cuts came in the portion of the CEQ budget allocated for research, a tacit rebuke of the kind of work the agency had undertaken under

Carter for it’s comprehensive 1980 systems-based Global 2000 Report to the President

(or Global 2000, as it was called)—a report, not coincidentally, that dealt in part with the potential environmental impacts of climate change.554

The transition had a profound impact on DoE and CEQ, but the most notorious bureaucratic changes initiated in what Sam Hays refers to as the “Reagan

Antienvironmental Revolution” occurred within those agencies most responsible for the everyday management of the environment. During the 1980 campaign, Reagan had mobilized moderate Republican conservationists like Russell Train and William

Ruckelshaus—both former EPA administrators—to help him establish a positive environmental image. Once elected, though, he brought in a new group of highly conservative, pro-business, pro-industry advisors tied to the Washington, D.C. based

Heritage Foundation to oversee a reshaping of American environmental policy. The group sought to control policy by appointing loyal right-wing political acolytes as policymakers. Reagan tapped Ann Gorsuch, a leader of the Colorado legislature’s

Republican Right, a lawyer for clients in both mining and agriculture, and an open opponent of federal regulation, to head up the EPA. Gorsuch, in turn, named 15 like- minded subordinates, 11 of whom had ties to the very industries the EPA was supposed

553 Tom Delaney, “Memorandum for Al Hill,” March 12, 1981, James A. Baker, III Files, Council on Environmental Quality, Folder 2, Box 1. See also Lance Gay, “Environmental Programs Facing Deep Budget Cuts,” Washington Star, March 11, 1981. 554 Gerald O. Barney, Global 2000 Report to the President: Entering the 21st Century, (Washington, D.C.: U.S. Government Printing Office, 1980).

240 to regulate.555 John Crowell, a lawyer for the world’s largest buyer of National Forest timber, took charge of the U.S. Forest Service, and Robert Harris, the leader of a legal effort to overturn the Surface Mining Control and Reclamation Act of 1977, became the head of the Office of Surface Mining, the very agency he had been fighting.556

The Antienvironmental Revolution’s coup de grace came in the person of Reagan’s new Secretary of the Interior, James G. Watt.557 An evangelical Christian and the founder of the Mountain States Legal Foundation—a firm dedicated to “limited and ethical government” and “economic freedom”—Watt openly opposed American conservation efforts, both as a matter of policy and as a matter of faith.558 He described environmentalism as “a left-wing cult dedicated to bringing down the type of government

I believe in.” “My responsibility,” he told The Wall Street Journal in 1981, “is to follow the Scriptures which call upon us to occupy the land until Jesus returns.”559 Watt committed the Department of Interior to “mine more, drill more, cut more timber,” and from his first day in office, he made it his mission to relax environmental regulations, to promote natural resource exploitation, and to reduce federal land ownership to the greatest extent possible. During his first year in office, he failed to list a single new

555 Kirkpatrick Sale, The Green Revolution: The American Environmental Movement, 1962-1992 (New York: Hill and Wang, 1993): 51; Hays, Beauty, Health, and Permanence, 494; Robert Gottlieb, Forcing the Spring: The Transformation of the American Environmental Movement (Washington, D.C.: Island Press, 1993): 292-296. 556 Hays, Beauty, Health, and Permanence, 494; Sale, The Green Revolution, 50. See also Jonathan Lash, A Season of Spoils: The Reagan Administration’s Attack on the Environment (New York: Pantheon Books, 1984) 557 The idea that Reagan had launched an anti-environmental revolution pervaded the media of the time, and is not just a product of Hays’s historical analysis. For example, in an article meant to vindicate Watt—but one that, in hindsight, falls comically short—conservative columnist George Will called Watt the “Robespierre of Western Resistance,” at once a nod to the power of Watt’s position and to the reactionary character of the Reagan Administration’s environmental policy. George Will, “The Robespierre of Western Resistance,” The Washington Post, January 15, 1981. 558 Mountain States Legal Foundation, http://www.mountainstateslegal.org/index.cfm. 559 The quotation appears in and is most frequently cited as “James Watt & The Puritan Ethic,” The Washington Post, May 24, 1981, but the Post article attributes the quotation to The Wall Street Journal (with no date).

241 endangered species—a record period of inaction (376 days) broken only recently by

Secretary of the Interior Dirk Kempthorne of the George W. Bush Administration.560 As incoming Secretary of the Interior, he vowed to “strike a balance” between the interests of environmentalists, recreationalists, business, and industry, but for environmentalists like Russell Peterson of the Audubon Society, this “balance” was a farce. As Peterson put it, “to put him in charge of the Interior Department is a crime.”561

The Reaction to the Reaction: Scientists, Environmentalists, and Democrats

The Reagan Antienvironmental Revolution represented the most staggering and comprehensive peacetime rollback of environmental policy in the history of the conservation movement. But it also helped to unite and energize a new coalition of liberal, pro-environment, anti-Reagan dissenters—a coalition that paired professional environmentalists with a wide variety of scientists, including many climate scientists, as

560 A former Governor of Idaho, Kempthorne received recognition for his feat from the Center for Biological Diversity in the form of the “Rubber Dodo Award,” a prize presented annually to “a deserving individual in public or private service who has done the most to drive endangered species extinct.” Watt’s inaction on endangered species led to an amendment to the ESA meant to counter the strategy of controlling policy by controlling policy-makers. The amendment set firm timelines and specific guidelines for listing endangered species, as well as statutory penalties for violating deadlines. Center for Biological Diversity, “Secretary of the Interior Dirk Kempthorne Wins 2007 Rubber Dodo Award,” Press Release, August 24, 2007, http://www.biologicaldiversity.org/news/press_releases/kempthorne-08-24-2007.html. Alaska Governor Sarah Palin won the award in 2008; land speculator Michael Winer was given the award in 2009. 561 Ironically, while Watt proved himself a walking catastrophe for the environment itself, his objectionable statements and actions helped kindle a revival of America’s environmental movement. In the midst of their political and policy defeats at the hands of Watt and the Reagan Administration, environmental organizations of all stripes—alongside traditional liberal interest groups like the American Civil Liberties Union and the National Organization for Women—saw their memberships mushroom. As a group, these organizations saw a 33% increase in fundraising returns in 1981, in large part as a result of the combination of direct mail campaigns and public coverage of Watt’s nomination hearings. James M. Perry, “Liberals Find Use for Reagan Crowd—As a Rallying Cry,” The Wall Street Journal, March 30, 1981; Andy Pasztor, “Watt Tells Senate He Would Balance Use, Protection of Resources as Interior Chief,” The Wall Street Journal, January 8, 1981. See also U.S. Congress, House Committee on Interior and Insular Affairs, Hearing, Briefing by the Secretary of the Interior, 97th Congress, 1st Session, February 5, 1981 (Washington, D.C.: U.S. Government Printing Office, 1981); U.S. Congress, Senate Committee on Energy and Natural Resources, Hearings, James G. Watt Nomination, 97th Congress, 1st Session, January 7 and 8, 1981 (Washington, D.C.: U.S. Government Printing Office, 1981).

242 well as with a vocal minority of Congressional Democrats. Already, Global 2000 and the associated issues of acid rain, ozone depletion, and deforestation had sparked a hesitant cooperation between these groups in dealing with global environmental problems. It was their common response to the Reagan Reaction, however, that brought these groups together to make global warming into a mainstream, Democratic environmental issue in the 1980s.

Political dissent in the early 1980s took many forms, but especially for

Congressional Democrats, the first line of defense came in the familiar form of

Congressional hearings.562 In the wake of Reagan’s resounding electoral victory and his subsequent whirlwind of deregulation activity, Democrats hoped to curb the new

Administration’s political momentum and set parameters on the President’s popular mandate by publicly challenging Reagan’s appointed officials and drawing out the implications of the Administration’s budget requests. For the most part, it was a form of politics as usual. As Paul Tsongas of Massachusetts noted in a January, 1981 hearing on the nomination of James Watt, few members of Congress realistically expected to block

Reagan’s appointments or significantly alter his Administration’s overall approach to environmental regulation.563 Rather, they sought to extract as high a political price as possible for their eventual approval of Reagan’s initiatives by undercutting the credibility

562 See, for example, U.S. Congress, James G. Watt Nomination; U.S. Congress, House Committee on Science and Technology, Hearing, Environmental Protection Agency Research and Development Posture, 97th Congress, 1st Session, October 22, 1981 (Washington, D.C.: U.S. Government Printing Office, 1981); U.S. Congress, House Committee on Appropriations, Hearings, Department of the Interior and Related Agencies Appropriations for 1982, Part 5, 97th Congress, 1st Session, Feb. 25, Mar. 4, 6, 10-12, 16, 20, Apr. 7, 1981, (Washington, D.C.: U.S. Government Printing Office, 1981). 563 “Being confirmed is a piece of cake,” Tsongas told Watt, “You will be confirmed…That is not my doing. I am speaking mathematically.” U.S. Congress, James G. Watt Nomination, 74.

243 of his administrators and questioning the wisdom of his policies.564 Technically, whomever Reagan appointed to administer federal agencies served not only the President, but also Congress. Much as Edmund Muskee and William Proxmire had used

Congressional hearings to call attention to the legally dubious expansion of executive power under President Nixon, in the early 1980s Democrats used hearings as a way to reign in Reagan’s aggressive policymakers by publicly reminding them of their

Congressional dependence.565

The hearings also provided a forum for professional environmentalists to voice their opposition to Reagan’s appointments and environmental policies, a complement to their efforts to mobilize their constituents against the new President through the popular press and direct mail campaigns. Their testimony spoke to the ongoing realignment of the politically fragmented American environmental movement with the mainstream

Democratic left. The American conservation movement had deep Republican roots, but as David Brower of Friends of the Earth and Russell Peterson of the typically more conservative National Audubon Society intimated in the Watt hearings, the new

Administration’s policies threatened to accelerate a recent trend away from the

Republican Party among the conservation movement’s rank and file.566 Meanwhile, as

564 In Watt’s case, Democrats focused on the conflict of interest involved in cases where the Secretary of the Interior had to decide upon how to enforce certain laws when his foundation, the Mountain States Legal Foundation, brought suit challenging the prosecution of those laws. Both Henry Jackson of Washington and Dale Bumpers of Arkansas got extensive mileage out of the conflict of interest question in Watt’s nomination hearing, although neither had any intention of seriously trying to block Watt’s confirmation. U.S. Congress, James G. Watt Nomination, 49-51, 55, 59-61. 565 In February of 1982, the House Committee on Appropriations called a hearing to deal specifically with Watt’s failure to submit to Congressional oversight. U.S. Congress, House Committee on Appropriations, Hearing, Secretary Watt’s Refusal to Cooperate with Congress, 97th Congress, 2nd Session, February 24, 1982 (Washington, D.C.: U.S. Government Printing Office, 1982). 566 Peterson was actually a Republican Governor of Delaware in the late 1960s and early 1970s, and only switched his party affiliation in 1996, though he began to support Democratic political candidates as early as the 1980s, largely because of his experiences with the Reagan Administration. Both he and Brower,

244 leaders within more liberal organizations like the Sierra Club and NRDC saw their few

Republican conservationists allies—allies like Russell Train and Bill

Ruckelshaus—increasingly marginalized in favor of administrators like Watt and

Gorsuch, they too moved steadily to the left.567

The confrontational Watt hearings set the tone for Democrats’ Congressional response to the Reagan Antienvironmental Revolution, but the Administration’s energy policy provided the most important locus of discontent among Reagan’s political, scientific, and environmental critics. Reagan’s popular commitments to national energy independence and an end to federal energy pricing controls relied on a number of policies that ran counter to conservation goals. These included increased domestic oil and mineral extraction, more domestic coal-fired power plants and environmentally destructive synfuels production, an increase in domestic nuclear energy production, and the broader deregulation of the energy industry. The President proposed to rely on private enterprise rather than government money to fund the vast majority of future energy research. He appointed a dentist as the Secretary of Energy and he later stripped

Carter’s already-paid-for solar hot water panels from the White House roof.568 When it came to energy, there was something for everyone to loath about Ronald Reagan.

Reagan’s semi-secret budgetary attack on the Department of Energy provided his opponents with an especially potent opportunity to publicly attack the President, both for his energy policy and for his broader assault on science and technology funding. Led in

along with Brock Evans of the Sierra Club and a number of other environmentalists, testified in opposition to Watt’s nomination, though to no avail. See U.S. Congress, James G. Watt Nomination, 153-202. 567 Gus Speth interview, 6/15/09. 568 Jim Edwards hardly garnered the type of press that James Watt did, but George Woodwell recalls that the scientific community was incensed by Edwards’ appointment. As I conducted interviews with Woodwell, Schneider, Hayes, and others, I found that the mention of Edwards’ name still elicits a reaction of anger and disgust.

245 the House by Don Fuqua of Florida, James Scheuer of New York, Harold Volkmer of

Missouri, George Brown of California, and Albert Gore of Tennessee, Democrats focused on three main problems with Reagan’s plan to restructure the DoE and cut the agency’s funding. First, they sought to demonstrate that dismantling the DoE would not in the end save taxpayers money, as the Reagan Administration claimed; rather, the agency’s destruction would cost the government money in the long run. Second,

Democrats went to great lengths to show how a breakup of the DoE would leave the government utterly unable to monitor, regulate, conduct research on, or even effectively oversee the production and consumption of energy in the United States, leaving the energy sector open to the dangerous and irresponsible boom and bust cycle of an unrestricted and easily corruptible free market. Finally, Al Gore, George Brown, and

James Scheuer took the lead in emphasizing the potential environmental, economic, and national defense consequences of dismantling the DoE’s research and development

budget. It was in this context that Gore and his colleagues reintroduced the issue of CO2- induced global warming.

A year earlier, in April of 1980, Congressmen George Brown, James Scheuer, and

Paul Tsongas had held hearings before a House Committee on Energy and Natural

Resources on the “Effects of Carbon Dioxide Buildup in the Atmosphere.”569 The hearings, which included (among others) NRDC co-founder and former CEQ chair Gus

Speth, biologist George M. Woodwell of the Brookhaven National Laboratory, Wally

Broecker of The Lamont Observatory, NCAR’s William Kellogg, and former CEQ member Gordon MacDonald, then of the Mitre Corporation, provided a sneak peek into

569 U.S. Congress, House Committee on Energy and Natural Resources, Hearing, Effects of Carbon Dioxide Buildup in the Atmosphere, 96th Congress, 2nd Session, April 3, 1980 (Washington, D.C.: U.S. Government Printing Office, 1980).

246 the climatic conclusions of the forthcoming Global 2000 Report. Politically, Brown,

Scheuer, and Tsongas hoped to use CO2-induced warming as an example of the long-term environmental consequences of a Carter energy policy then tilting away from conservation and renewables and toward coal and synfuels.570 But for the most part the hearing was soft on politics and proceeded amicably, the only tension a product of an ongoing scientific controversy between Woodwell and Broecker. MacDonald’s policy

recommendations—that CO2 research be incorporated into energy research, that the

President reconsider renewables, that the U.S. take the lead in international CO2 research, and that more money be made available for this type of research—closely resembled the type of recommendations that scientists and members of Congress had made since the mid-1970s. The hearing ruffled few feathers.

When Gore and Scheuer called a hearing before the House Committee on Science and Technology on “Carbon Dioxide and Climate: The Greenhouse Effect” in 1981,

however, the tone of the CO2 discussion took what one scientists later described as an

“ugly turn.”571 The hearings began benignly enough, with relatively uncontroversial testimony from Roger Revelle, Joseph Smagorisnki, Stephen Schneider, and Lester Lave, each of whom described their concerns about the potential dangers of global warming and outlined the specific areas of research that could help scientists provide the type of

572 information Congress would need to make effective policy on CO2 and climate. In part, the proceedings reflected Gore’s and Scheuer’s genuine, continued concern over an increasingly disconcerting environmental issue—an issue that Gore would make a

570 U.S. Congress, Effects of Carbon Dioxide, 1980, 3. 571 Schneider, Contact Sport, 88. U.S. Congress, House Committee on Science and Technology, Hearing, Carbon Dioxide and Climate: The Greenhouse Efffect, 97th Congress, 1st Session, July 31, 1981 (Washington, D.C.: U.S. Government Printing Office, 1981). 572 U.S. Congress, Carbon Dioxide Buildup, 1981, 1-82.

247 centerpiece of his long political career.573 But the Congressmen also mobilized these scientists to demonstrate the Reagan Administration’s myopia and even negligence in

cutting government funding for CO2 research at the DoE. Reagan’s aides had kept the details of their DoE budget reallocation secret, but Gore and Scheuer—along with the scientists who they invited to testify—knew that Administration officials planned to cut

CO2 research along with the majority of DoE’s other environmental research programs.

In the second half of the hearing, Gore and Scheuer called on Reagan’s new appointments in the Department of Energy, the Office of Science and Technology Policy, and the National Climate Policy program, and challenged them to defend their yet-to-be- released budget cuts, not only in the face of the facts of state of the art climate science, but quite literally in the face of the scientists who had produced it.574

The Democrats’ public approach to CO2 yielded three important results. First,

Gore’s 1981 hearings on CO2, along with subsequent CO2 hearings in 1982 and 1984, helped to establish both personal and ideological relationships between prominent climate scientists (many of them already relatively liberal) and the mainstream political left.575

573 In March of 1981, Scheuer and Brown also led hearings on the oversight of the National Climate Program, a program increasingly tied to the U.S. Department of Agriculture and more and more distant from the Department of Energy under Reagan. Ironically, it was research conducted under the auspices of the USDA that allowed Reagan and his successor, George Herbert Walker Bush, to re-engage with the issue of global warming at a Presidential level in the late 1980s. See U.S. Congress, House Committee on Science and Technology, Oversight of the National Climate Program, Hearing, 97th Congress, 1st and 2nd Sessions, March 5, 1981 and March 16, 1982 (Washington, D.C.: U.S. Government Printing Office, 1982). 574 Stephen Schneider recalls the change in tone in Congress in the early 1980s, and he attributes the transformation to the introduction of Reagan appointees like Frederick Koomanoff and N. Douglas Pewitt to the conversation. And to some extent, the attribution is valid—Reagan officials did, as a point of policy, eliminate social and environmental science research from the federal budget, including much climate science research, and in their actions they also challenged the validity of the science they hoped to cut. But

Schneider fails to recognize that Democrats, too, sought to politicize the CO2 issue as a way to undercut Reagan, and their aggressive posture in the House in the early 1980s helped to polarize the increasingly partisan discussion in the years to come. Schneider, Contact Sport, 88. 575 U.S. Congress, Carbon Dioxide and Climate; U.S. Congress, House Committee on Science and Technology, Carbon Dioxide and the Greenhouse Effect, Hearing, 98th Congress, 2nd Session, February 28, 1984 (Washington, D.C.: U.S. Government Printing Office, 1984).

248 Democrats had sponsored hearings on CO2 and climate in the past, but unlike in the informational hearings of the1970s, the confrontational nature of the hearings of the early

1980s all but forced scientists to take sides. Reagan’s appointees—in particular N.

Douglas Pewitt, an assistant to Energy Secretary Jim Edwards—alienated scientists like

Revelle and Schneider by threatening to cut their DoE funding. In the process of defending the cuts, these appointees directly challenged both the value and substance of scientists’ work.576 Gore, Brown, and Scheuer provided an attractive alternative. Most obviously, the Democrats called hearings to give scientists a voice in trying to protect

funding for CO2 research; they were natural allies. But Brown, Scheuer, and even more notably Gore, a one-time student of Revelle’s at Harvard, also demonstrated a high level of scientific literacy and a genuine concern over the issue of global warming that impressed the scientists on a personal level. These personal relationships would continue to develop as Gore climbed through the political ranks over the next two decades.

Second, in this and other hearings on CO2 and on the future of the DoE more

broadly, Gore helped to publicly establish CO2-induced climate change as an important

577 issue associated with the ongoing debate over energy. “If we take CO2 seriously,”

Lester Lave of the Brookings Institution told The New York Times for an article about the hearing in July of 1981, “we would change drastically the energy policy we are

578 pursuing.” The 1981 CO2 hearing failed to save the agency’s carbon dioxide program

576 Schneider, Contact Sport, 88-92; U.S. Congress, Carbon Dioxide and Climate, 1981. 577 U.S. Congress, Carbon Dioxide and Climate; U.S. Congress, Carbon Dioxide and the Greenhouse Effect. 578 The Times article more specifically targeted the development of synthetic fuels from coal and shale, criticizing the Administration’s $20 billion budget for synfuel development. “Synthetic Fuels Called a Peril to the Atmosphere,” The New York Times, July 31, 1981. In his testimony before the House

Committee on Science and Technology in 1981, Lave very thoughtfully discussed the role of CO2 not just as a specific problem but as a symbol for a whole category of problems that transcended the chronology of the political cycle. He chastised the shortsightedness of elected officials and framed the CO2 case as a sort

249 from the budgetary meat axe, but the press picked up the energy story and ran with it.

“The threat of an overload of carbon dioxide, changing the planet’s climate, is perhaps not immediate,” read a Washington Post Article urging the President to pay more attention to a DoE Energy Research Advisory Board report in December of 1981. “But it would impose an absolute limit on fuel consumption, making the oil crisis and dislocations of the past seem trivial.”579 Since the late 1970s, scientists like MacDonald

had argued that CO2 should be incorporated into energy research; in the 1980s, Gore and his Democratic colleagues ensured that CO2 had a place in energy politics.

Finally, in part because of the focus on energy, Gore’s hearings all but demanded that American environmental organizations enter the public discussion about global

warming. Environmentalists took heed. “There is of course no way that researching CO2 build-up cannot call into question the current Administration’s energy policy,” wrote

Anthony Scoville in Friends of the Earth’s newsletter, Not Man Apart, in 1982. “The

Administration is cutting funding into CO2 research so that it can justify its deliberate

580 ignorance of the issues at hand.” In testimony submitted to Gore’s 1982 CO2 hearing,

FoE’s Rafe Pomerance argued that “global warming could be one of the most serious and irreversible environmental problems yet faced by man.”581 Echoing both scientists and

Democratic Congressmen, Pomerance outlined Friends of the Earth’s new position:

“…climate change must now be included as part of energy and economic

of vanguard of progressive thinking. U.S. Congress, Carbon Dioxide Buildup, 1981, 60-67. See also David

M. Burns, “Climate and CO2,” The New York Times, April 17, 1981. 579 “Good Advice on Energy,” The Washington Post, December 28, 1981. Journalists also made a direct connection between the CO2 problem and Democrats’ and environmentalists’ continuing criticism of the synfuels program initiated under Carter. See “Synthetic Fuels Called a Peril to the Atmosphere,” The New York Times, August 1, 1981. 580 Anthony Scoville, “Why the U.S. Ignores the Greenhouse Effect,” Not Man Apart, pg 24, reprint in U.S Congress, “Carbon Dioxide and Climate,” 1982, 147. 581 U.S Congress, Carbon Dioxide and Climate, 1982, 146.

250 innovation policy. Faced with the possible threat of global warming it is

irresponsible to decimate programs for energy conservation and solar energy as the

Reagan Administration is doing.”582

The Environmental Defense Fund, though it did not participate in the hearings, soon decided that the issue was sufficiently important to hire Harvard atmospheric physicist

583 Michael Oppenheimer as a senior scientist. NRDC continued to work CO2 into its international policy positions, and Gus Speth’s new organization, the World Resources

Institute, made climate change one of its central foci. It would take until the second half

of the decade for environmentalists to meaningfully coordinate their efforts on CO2, but

Gore’s hearings ensured that when global warming finally began to stick as a mainstream issue in the summer of 1988, environmental organizations—and their constituents—were ready to act.

Climate Change and the Reagan Bureaucracy

The fight over funding for CO2 research extended beyond the halls of Congress; government scientists also fought to maintain control over climate research within the bureaucracy itself. Here the results were mixed. Individual government scientists used

their government-sponsored research on CO2 as a way to highlight the folly not only of conservative science, but also of conservative energy policy and conservative science

policy—especially vis-à-vis funding for CO2 research. Reagan appointees in turn used their control over science budgets and administrative priorities to privilege a scientifically

582 Ibid. 583 “Michael Oppenheimer CV,” available at Princeton University Program in Science, Technology, and Environmental Policy, http://www.princeton.edu/step/people/faculty/michael-oppenheimer/. See also Michael Oppenheimer, “To Delay on Global Warming,” The New York Times, November 9, 1983.

251 conservative position that in turn supported a conservative approach to energy CO2.

Originally a battle over science funding and energy policy, the political battle over CO2 soon expanded to include a struggle over the conclusions of climate science itself.

The first important conflict between the Administration and its own scientists occurred at NASA’s Goddard Institute for Space Studies (GISS), headed by James

Hansen, and again it involved a dispute over funding from Reagan’s Department of

Energy. A taciturn Iowan and the picture of a careful, plodding scientist, Hansen and a group of planetary climate modelers at NASA had spent much of the late 1970s developing a new kind of General Circulation Model designed to reveal the long-term, average characteristics of climate independent of the short-term, amplitudinal characteristics of weather.584 Funded by an internal NASA grant under the Carter

Administration, Hansen’s group used the model—called “Model Zero” until 1981, when

GISS renamed it “Model II”—to explore the Earth’s climatic response to a doubling of

atmospheric CO2 given a variety of different feedback processes and systemic sensitivities. Unlike other, higher resolution GCMs, the course model also allowed

Hansen and his colleagues to bracket the potential climatic impacts of the oceans by running the model multiple times under different assumptions about ocean mixing and

584 Initially, under the direction of Robert Jastrow, the GISS group focused primarily on developing better short-term, high-resolution models of atmospheric circulation designed to help forecasters make better weather predictions. In the mid-1970s, however, Jastrow began to distance himself from the institute, leaving program directors like Hansen—then overseeing planetary and climatic research at the institute—free to pursue other aspects of atmospheric modeling. Interview with Jim Hansen by Spencer Weart at Goddard Institute for Space Studies, New York City, October 23, 2000, Niels Bohr Library & Archives, American Institute of Physics, College Park, MD USA, http://www.aip.org/history/ohilist/24309_1.html. For more on NASA’s shift to Earth studies, and on the subsequent Earth Observation System, see “NASA Atmospheric Research in Transition,” “The Quest for a Climate Observing System,” and “Mission to Planet Earth: Architectural Warfare” in Erik Conway, Atmospheric Science at NASA: A History (Baltimore: Johns Hopkins Press, 2008), 122-153, 198-242, 243- 275.

252 heat capacity.585 Using historical climate data from as early as 1880, the model consistently produced results “roughly consistent” with measured reality.586

When Hansen finally published the Model II results in Science in 1981, his conclusions did not go over well with an Administration wary of the Global 2000 results

and bent on downplaying the CO2 issue. “The global warming projected for next century is of almost unprecedented magnitude,” Hansen and his colleagues wrote.587 The warming could lead, among other things, to the relatively rapid collapse of the West

Antarctic Ice Sheet, producing an overall sea level rise of up to six meters—enough to

“flood 25% of Florida and Louisiana, 10% of New Jersey, and many other lowlands throughout the world.”588 The paper also made specific reference to the need for a more

diversified national energy portfolio in light of the potential climatic impacts of CO2 from fossil fuels. “The degree of warming will depend strongly on the energy growth rate and choice of fuels for the next century” the GISS scientists wrote. Their recommendations aligned with those of Carter’s CEQ; they ran directly counter to the new President’s energy policy:

“CO2 effects on climate may make full exploitation of coal resources undesirable.

An appropriate strategy may be to encourage energy conservation and develop

alternative energy sources, while using fossil fuels as necessary during the next few

585 Model Zero’s grid used cells on the order of a thousand square kilometers—huge by the standards of weather modelers, but adequate for studies of climate. Under Hansen, GISS stood at the cutting edge of climate modeling; during the National Academy of Science’s preparation of its seminal 1979 report on CO2 and climate, committee chair Jule Charney consulted with Hansen repeatedly in order to incorporate the GISS results into the NAS survey of climate research. Interview with Hansen by Weart, October 23, 2003. See also Conway, Atmospheric Science at NASA, 202. 586 Ibid. See also Walter Sullivan, “Study Finds Warming Trend That Could Raise Sea Levels,” The New York Times, August, 22, 1981. 587 J. Hansen et al, “Climate Impact of Increasing Carbon Dioxide,” Science, vol. 213, no. 4511 (August 28, 1981): 966. Also cited in Conway, Atmospheric Science at NASA, 202. 588 Hansen et al, “Climate Impact,” 1981, 965.

253 decades.”589

Worse still for the new Administration, Hansen gave a copy of the paper to science journalist Walter Sullivan.590 For Science, Model II represented the cutting edge of the continually improving world of GCMs; at The New York Times, a six meter sea level rise made front page news.591

The Administration responded quietly and predictably; they downplayed the newspaper report and went after Hansen’s funding. Hansen’s 1977 NASA grant terminated on schedule during the Carter-Reagan transition. In 1980, amidst uncertainty over the future of the GISS climate program, David Slade, then still in his role as the

DoE’s chief CO2 man, committed the DoE to provide support for Hansen’s work within

the context of the AAAS/DoE CO2 program moving forward. After Slade’s departure, however, Hansen had to appeal to Frederick Koomanoff—the same Fred Koomanoff called by Gore and Scheuer to testify on behalf of the Reagan administration’s cuts in

CO2 research at DoE—to secure DoE funding. Hansen’s model had drawn criticism from a number of quarters within the scientific community, and Koomanoff used these criticisms as grounds for denying Hansen’s application, effectively curtailing the GISS project.592 The subsequent funding cuts forced Hansen to lay off five researchers.593

589 Ibid., 966. 590 Hansen became acquainted with Sullivan over meals at Robert Jastrow’s favorite lunch spot, a Chinese restaurant called the Moon Palace across the street from the GISS labs. Weart interview of Hansen, 10/23/2000. 591 Sullivan, “Study Finds Warming Trend,” 8/22/81. 592 Stephen Schneider, Ishtiaque Rasool, and Michael MacCracken provided perhaps the most direct scientific criticism of Hansen’s Model II—somewhat surprising considering that Schneider, a liberal, progressive climate scientist convinced of the CO2 threat, had worked with Hansen briefly as a graduate student in 1970, and Rasool, another GISS scientists, had helped to get Hansen his job at NASA in the first place. See Spencer Weart Interview of James Hansen, 10/23/2000. See Michael C. MacCracken, S. B. Idso, J. Hansen, D. Johnson, A. Lacis, S. Lebedeff, P. Lee, D. Rind and G. Russell, “Climatic Effects o Atmospheric Carbon Dioxide [exchange of letters],” Science vol. 220, no 4599 (May 20, 1983): 873-875; S. Ichtiaque Rasool et al., "On Predicting Calamities [with Reply]," Climatic Change vol. 5, no. 2 (June, 1983): 201-04. See also Stephen H. Schneider, William W. Kellogg, V. Ramanathan, Conway B. Leovy

254 More insidiously, as Hansen told Erik Conway in 2006, the new DoE officials also warned him not to share Model II with would be collaborators at Pennsylvania State

University, suggesting that otherwise these researchers, too, might lose their government support.594

Silencing Hansen represented the first in a series of ad hoc steps by the Reagan

Administration to use the mechanisms of the government bureaucracy to quash the fruits

of Carter-era research on CO2 and climate change. The effort included more than an overall reduction in funding for CO2 research and the bureaucratic marginalization of individual scientists; it also involved capitalizing on both professional and political divisions within the scientific community in order to manage the public scientific

message on CO2. The Administration used these tactics perhaps most effectively in handling two government-sponsored reports on CO2 and climate change in October of

1983, one produced by the National Academy of Sciences and one by the EPA.

By 1980, the climate science community had come to agree on a few basic

principles vis-à-vis CO2 and climate, but scientists from different disciplines and institutions continued to disagree over the timing and magnitude of a potential CO2-

induced warming. In 1979, a special National Academy of Sciences study group on CO2 and climate released a report that provided the first self-conscious scientific consensus on

and Sherwood B. Idso, “Carbon Dioxide and Climate [exchange of letters],” Science vol. 210, no. 4465 (October 3, 1980): 6-8. 593 Hansen talked at length about his presentation to Koomanoff in his 2000 interview with Spencer Weart. See also Weart, Discovery, 143-144; Conway, Atmospheric Science at NASA, 204. For a popular and polemical account of Hansen’s relationship with government officials in the 1980s, see Mark Bowen, Censoring Science: Inside the Political Attack on Dr. James Hansen and the Truth of Global Warming (New York: Plume, 2008). 594 Conway, Atmospheric Science at NASA, 204.

255 595 CO2 and warming. The group concluded that a doubling of CO2 would lead to a 1.5-

4ºC increase in the global mean temperature—a warming that might have significant social and environmental impacts.596 As scientific debate over Hansen’s 1981 paper showed, however, unanswered questions about the behavior of the oceans and the biosphere within the overall climate system made it difficult to predict how quickly and

how severely the climate might change in response to CO2 in the real world. Though both the NAS study and Hansen’s paper identified a moderate increase in temperature

associated with CO2 build-up in the recent past, the models revealed only the faintest

signal of a climatic response to the nearly 15% increase in CO2 that had occurred since the mid-19th century.597 Whereas Hansen argued that the Earth would begin to feel the impacts of global warming early in the 21st century, a number of other scientists

contended that the ocean-induced lag in the CO2-climate system could last as long as a century or more, while still others all but discounted the lag altogether.598

The 1983 NAS and EPA reports transformed these scientific disagreements into policy positions. Chaired by Scripps Institution of Oceanography director William

Nierenberg, the National Academy of Sciences study arose out of the National Energy

595 See chapter 4. See also Jule Charney et al., Carbon Dioxide and Climate: A Scientific Assessment, National Research Council, Ad Hoc Study Group on Carbon Dioxide and Climate (Washington, D.C.: National Academy Press, 1979). 596 The group actually stated that there was “no reason to believe that these changes will be negligible,” an odd double-negative and the sort of grammatical equivocation that would come to characterize later reports on global warming. Charney, Carbon Dioxide and Climate, vii. 597 Hansen estimated that CO2 stood at 293ppm in 1880, 335ppm in 1980, and would reach 373ppm by 2000. Despite a Northern Hemisphere cooling between 1940 and 1970, NASA’s dataset showed an overall increase of 0.4ºC during the twentieth century, a figure “roughly consistent” with Model II results. Hansen et al, “Climate Impact,” 1981, 960; Conway, Atmospheric Science at NASA, 202. 598 Even here, subdivisions existed between scientists over the meaning of an ocean lag. Hansen argued that the oceans did indeed constitute a mid-term climate stabilizer—in fact, he and Wally Broecker agreed that the oceans masked the climate signal of CO2. For both Broecker and Hansen, this masking looked ominous; when the lag time was up, how much stored CO2and latent heat would the ocean release, and how fast? Conway, Atmospheric Science at NASA, 205; See also Weart interview with James Hansen, 10/23/2000; Weart interview with Wallace Broecker, 11/14/1997.

256 Security Act of 1980. Following the push during the Carter Administration to

incorporate CO2 research into federal energy policy, the Act mandated that federal officials work with the National Academy of Sciences to produce a follow-up to the 1979

“state of the science” report. 599 The new report specifically addressed the relationship between the fossil fuel energy mix and the broader social and economic impacts of a

600 CO2-induced warming. It also broadened the 1979 assessment to more fully include the biosphere and the oceans in the overall climatic picture.

The Nierenberg committee noted that the potential impacts of atmospheric CO2 represented a “cause for concern,” but overall the report downplayed the urgency of the issue and recommended a wait, do more research, and see approach to the problem.601

Thomas Schelling of Harvard and William Nordhaus of Yale, the committee’s two economists, argued that in a world of slow and gradual climatic change, humans would

602 likely learn to adapt as fast as the CO2 problem developed. Tied to energy, pollution,

603 and the global environment, the report said, CO2 presented an intractable problem. But as Nierenberg told The New York Times, the committee felt that “we have 20 years to examine options before we make drastic plans. In that 20 years we can close critical gaps

599 Jay Keyworth, “Memorandum for Ed Meese,” October 23, 1983, Environmental Protection Agency Folder, Box 3, George A. Keyworth files, Archives, Ronald Reagan Presidential Library, Simi Valley, CA. See also Naomi Oreskes, Erik Conway, and Matthew Shindell, “From Chicken Little to Dr. Pangloss: William Nierenberg, Global Warming, and the Social Deconstruction of Scientific Knowledge,” Historical Studies in the Natural Sciences, vol. 38, no. 1 (Winter, 2008): 122. 600 Oreskes et al., “Chicken Little,” 122. 601 National Academy of Sciences, Carbon Dioxide Assessment Committee, Changing Climate (Washington, D.C.: National Academy of Sciences, 1983). See also Weart, Discovery, 146; Conway, Atmospheric Science at NASA, 205. 602 From Oreskes, “Chicken Little,” 124. Oreskes cites Thomas Schelling et al. to Philip Handler, 18 Apr 1980, “Ad hoc Study Panel on Economic and Social Aspects of Carbon Dioxide Increase,” on 11. Courtesy of Janice Goldblum, National Academy of Sciences Archive. 603 Phillip Shabecoff, “Haste on Global Warming Trend is Opposed,” The New York Times, October 21, 1983.

257 604 in our knowledge.” With good scientists on the case, CO2 was no reason to upset the economy through major changes in energy policy.

The 1983 study’s conservative conclusions clearly reflected the influence of conservative men within both the Academy and the Administration. There is little direct evidence that the Reagan Administration actively manipulated the Nierenberg report; with conservative scientists leading and funding the study, they didn’t need to. The

Energy Security Act mandated that the Office of Science and Technology Policy administer the NAS study, which gave George A. Keyworth, III, the President’s science advisor, control of the Nierenberg committee’s funding through 1983.605 Nierenberg himself, though selected as chair of the committee before the 1980 election because of his background as an experienced scientists and capable administrator, was a political conservative who later served on Reagan’s 1981 transition team, and as an advisor to the

President on atmospheric issues throughout his two terms in office.606 To the extent that the Administration put pressure on NAS group to toe the line, it came relatively subtly in the form of a budgetary threat from Fred Koomanoff. Komanoff held the purse strings

for the little DoE funding still available for CO2 research in the early 1980s, and he quietly made it known that he hoped the Academy’s recommendation would line up with the Administration’s low-key position on the issue.607 Ultimately, the fragmented, 496-

page, multi-author report expressed a variety of viewpoints on CO2, but Nierenberg controlled the report’s presentation and ultimately penned its summary. By framing the

604 Shabecoff, “Haste,” 10/21/83. 605 Keyworth memo for Meese, 10/23/83; Ronald B. Frankum to William S. Heffelfinger, October 6, 1982, Department of Energy 1981/1982, File 4 of 8, Box 3, George A. Keyworth Files, Ronald Reagan Presidential Library. 606 Oreskes et al., “Chicken Little,” 122. 607 Ibid., 136.

258 issue as a scientific problem of little immediate concern and subtly undercutting the consensus of the 1979 Charney study, Nierenberg gave the Administration the best report it could have hoped for.

The EPA report, by contrast, presented the CO2 situation as dire and acute.

Produced in a rush by Stephen Seidel and Dale Keyes as a preemptive response to the conservative NAS report, the EPA declared that “global greenhouse warming is neither trivial nor just a long-range problem.”608 Like the NAS scientists, Seidel and Keyes saw some warming as inevitable, and they focused on the need to develop strategies for adapting to climate change as well as for preventing it. But the EPA scientists’ fixation on the potentially disruptive impacts of climate change as early as the 1990s, along with the title of their study, Can We Delay a Greenhouse Warming?, also highlighted the costs

of a “wait and see” strategy that failed to account for CO2-induced warming in policies on energy and fossil fuels. Whereas the NAS committee saw no reason to overreact, Seidel and Keyes argued for a “soberness and sense of urgency” in response to global warming that could buy humanity time to adapt.609

Both reports essentially confirmed the inevitability of a CO2-induced warming, but not surprisingly, the Reagan Administration preferred the tempered conclusions of the

608 Technically, as the Administration pointed out in internal discussions about how to handle the report, the Seidel and Keyes study did not represent an official position of the EPA. Under the new EPA chief, William Ruckelshaus, the scientists had been given the liberty to investigate atmospheric issues on their own, and the report seems to have been the scientists’ own initiative. In the end, rather than risk revisiting the politically damaging shake-up that had occurred at the EPA after a series of scandals involving ideological appointments like Gorsuch and the her eventually incarcerated underling, Rita Lavelle (indicted on perjury chargers), Keyworth and Meese decided to let the EPA angle lie. Stephen Seidel and Dale Keyes, Can We Delay a Greenhouse Warming? (Washington, D.C.: Environmental Protection Agency, 2nd ed., 1983); Phillip Shabecoff, “EPA Report Says Earth Will Heat Up Beginning in 1990s,” The New York Times, October 18, 1983; “Key Points on Carbon Dioxide/Greenhouse Effect Report,” Environment—Carbon Dioxide Folder, OA11477, Danny Boggs Files, Ronald Reagan Presidential Library. For more on the EPA shakeup, see Hays, Beauty, Health, and Permanence, 513-520; Sale, The Green Revolution, 52; Lash, A Season of Spoils. 609 Seidel and Keys, Can We Delay?; Shabecoff, “EPA Report,” 10/18/83.

259 Nierenberg committee to the EPA’s alarming worst-case scenario. Keyworth and White

House Counsel Ed Meese decided to fight science with science.610 As Keyworth noted in a memo to Meese shortly before the release of the NAS report, the contrast between the two studies served the Administration well. “Although the Nierenberg report, like any other, has the potential for being misinterpreted,” Keyworth wrote, “I believe it is a sound report that should help defuse worst-case fears of the impacts of the greenhouse effect.”611

Keyworth used the cautious Nierenberg report to discredit the “unnecessarily alarmist”

EPA study in the press.612 Both reports made headlines, but by playing up the disparities between the studies, Keyworth managed to imbue the press coverage more with a sense of confusion and uncertainty than with one of concern.613 When the White House

followed up with an expert briefing on the CO2 issue for the President and the new head of the EPA, William Ruckelshaus, Keyworth predictably chose Nierenberg as their expert.614

610 By 1983, using scientific uncertainty and disagreements between scientists to defuse environmental concerns had become a tried and true tactic for the Reagan Administration. The CO2 case is particularly interesting, however, because whereas in other cases, the Administration could count on private sector scientists to help destroy consensus on environmental issues, with CO2 they had two quasi-governmental studies, both hewing pretty close to a pre-established NAS consensus, both of which the Administration had incentive to at least partially discredit. Despite lauding the NAS study, Keyworth’s contrasting of the two CO2 reports served to undermine the certainty of both reports—a novel approach for the anti- environmental OSTP. See Hays, “The Politics of Science” in Beauty, Health, and Permanence, 329-362. 611 Keyworth memo for Meese, 10/23/83. 612 Jay Keyworth, “OSTP Monthly Report for October 1983,” Memo for Ed Meese, October 28, 1983, OSTP Monthly Report, 1982-84, Folder 1 of 4, Box 6, George A. Keyworth, III Files. 613 See Arlen J. Large, “Warming of the Earth is Met with a Degree of Reassurance,” The Wall Street Journal, October 21, 1983; “How to Live in a Greenhouse,” The New York Times, October 23, 1983; Edgar Tasschdjian, “The Promise of the Greenhouse,” The New York Times, November 2, 1983; Michael Oppenheimer, “To Delay on Global Warming,” 11/9/83; Shabecoff, “EPA Report,” 10/18/83; Shabecoff, “Haste,” 10/21/83. 614 Keyworth memo for Meese, 10/23/83.

260 Nuclear Winter

The new politics of climate change did not revolve entirely around CO2 and fossil fuels; concerns over national defense and nuclear disarmament also helped to shape the political divisions within the climate science community in the 1980s, particularly as they related to the issue of “nuclear winter.”615 Coined by NASA’s Richard Turco in 1983, scientists used the term “nuclear winter” to describe a dramatic overall decrease in global mean temperatures that they believed would accompany a major nuclear exchange.

Nuclear winter had very little to do with CO2 and nothing to do with energy policy, but the issue brought together nearly every major player in atmospheric science in the 1980s, and the ensuing controversy helped to shape both the science and politics of global warming in the years to come. Public concerns about the climatic consequences of nuclear war, alongside interest from members of Congress and from the Defense

Department, created a new demand and a new source of funding for climate scientists’ expertise in atmospheric chemistry, atmospheric physics, and most importantly climate modeling. At the same time, the political salience of the issue provided these scientists with a public forum in which they could use their expertise to undermine the

615 Lawrence Badash provides a concise and well-narrated account of the nuclear winter saga from a political perspective in “Nuclear Winter: Scientists in the Political Arena,” Physics in Perspective, vol. 3, no. 1 (March, 2001): 76-105. Badash’s analysis addresses the episode in two important contexts. First, he looks at where the effort to use the issue of nuclear winter to influence nuclear defense policy fits into the chronologically broader story of physicists’ efforts to control nuclear weapons policy since the advent of the atomic bomb in 1945. Second, he investigates the role of the individual scientists involved—Carl Sagan in particular—in promulgating the “nuclear freeze” agenda of the anti-nuclear movement in the specific historical moment of the early 1980s. His is probably the best organized and most readable account of the issue, although a variety of other historians—mostly historians of atmospheric science and historians of physics—have tackled the issue from a variety of other perspectives. See also Lawrence Badash, A Nuclear Winter’s Tale: Science and Politics in the 1980s (Cambridge, MA: MIT Press, 2009). For nuclear winter in the context of atmospheric science, see Conway, Atmospheric Science at NASA, 206- 212; Weart, Discovery, 144-145; Schneider, Contact Sport, 95-108. For more on nuclear winter in the context of the anti-nuclear movement, see David S. Meyer, A Winter of Discontent: The Nuclear Freeze Movement and American Politics (New York: Praeger, 1990); Frances Fitzgerald, Way Out There in the Blue: Reagan, Star Wars, and the End of the Cold War (New York: Simon and Schuster, 2000), 179-82.

261 controversial national defense policies of an unfriendly Reagan Administration. The experience reinforced the relationships between outspoken climate scientists and both the mainstream political left and the American environmental movement, two communities closely (but often ambivalently) associated with disarmament efforts and the anti-nuclear movement. Because of these associations, however, the issue also alienated more politically conservative scientists, and the politically-charged scientific dispute the nuclear winter hypothesis engendered ultimately served to undermine climate scientists’ public credibility and added to the bitterness and rancor that had begun to characterize the

ongoing debate over CO2 and climate change.

The nuclear winter debate first arose out of an unlikely mix of politics, paleoclimatology, and ozone research. During the 1980 Presidential campaign, Reagan had stumped for a more robust and aggressive national defense strategy. He attacked

Carter’s halting attempts to cooperate with the Soviet Union on arms control as an attitude of “defeatism,” and he called the still unratified Strategic Arms Limitations

Treaty (SALT II) “fatally flawed.”616 He and his Secretary of Defense, Caspar

Weinberger, argued that Carter’s policies had opened up a “window of vulnerability” wherein the United States lacked the technological systems to respond to a Soviet first strike.617 The Administration called for “rearmament,” an expansion and updating of the

United States’ military arsenal and the means to deploy it.618 Weinberger argued that

616 Schaller, Reckoning, 129. 617 Richard Halloran, “Weinberger Expects 6-Month Delay For Renewal of Talks on Arms Curb,” The New York Times, January 7, 1981; Tom Wicker, “Beware of ‘Gaposis,’” The New York Times, January 9, 1981. 618 In 1981, the Administration unveiled a $180 billion plan to “revitalize the nation’s strategic deterrent.” The most notable and controversial aspects of the plan were the deployment of the new MX missile—a missile capable of carrying ten nuclear warheads and delivering them with a high level of accuracy—and the B-1 long range bomber, but Weinberger argued that communications systems that allowed for a flexible response to a nuclear threat and opened new strategy options represented an equally important part of the

262 rearmament would allow the U.S. to regain the nuclear parity at the heart of the policy of deterrence, but the Secretary of Defense also launched a series of studies on strategies for fighting—and winning—a protracted nuclear conflict with the Soviet Union.619

For foreign policy doves and American liberals already frightened by Reagan’s increasingly belligerent Cold War foreign policy and his push for rearmament,

Weinberger’s comments about “prevailing” in a nuclear conflict represented a divergence from the objectionable but still useful concept of Mutually Assured Destruction that underpinned deterrence. Their response included Congressional Hearings (led in part by

Al Gore, an impressively capable self-taught expert on nuclear defense), public exposés on nuclear destruction like Jonathan Schell’s 1982 The Fate of the Earth, and, most notably, a grassroots political effort called the Nuclear Freeze Movement that led a series of large anti-nuclear demonstrations and sponsored legislation—which passed the

House—resolving to cease the production of nuclear weapons.620 Foreign governments were equally horrified, and the already robust international disarmament movement also redoubled its efforts, both within national governments and at the United Nations.

In the United States and abroad, environmentalists stood in solidarity with this resurgent anti-nuclear movement. Since as early as the 1972 U.N. Conference on the

Human Environment—a conference that had its roots in Swedish concerns about nuclear

equation. See Richard Halloran, “Reagan Arms Policy Said to Rely Heavily on Communications,” The New York Times, October 12, 1981; Schaller, Reckoning, 127. 619 Halloran, “Pentagon Draws Up First Strategy for Fighting a Long Nuclear War,” The New York Times, May 30, 1982; Richard Halloran, “Weinberger Confirms New Strategy on Atom War,” The New York Times, June 4, 1982; Tom Wicker, “Crossing a Thin Line,” The New York Times, October 21, 1981. See also Caspar Weinberger, Fighting for Peace: Seven Critical Years in the Pentagon (New York: Warner Books, 1990). 620 Schaller, Reckoning, 128. Jonathan Schell, The Fate of the Earth (New York: Random House, 1982). See also Meyer, A Winter of Discontent; Thomas R. Rochon and David S. Mayer, Coalitions and Political Movements: The Lessons of the Nuclear Freeze (Boulder, CO: Lynne Rienner, 1997); Douglas C. Waller, Congress and the Nuclear Freeze: An Inside Look at the Politics of a Mass Movement (Amherst, MA: University of Massachusetts Press, 1987).

263 weapons and their dominance in international politics—environmental activists around the world had associated wars and nuclear armaments with environmental destruction and peace with conservation.621 Reagan’s defense policies reawakened longstanding concerns over the environmental impacts of military activities. In 1981, for example, a Sierra Club

Task Force on the Environmental Effects of Military Projects drafted a resolution recognizing the “unprecedented destruction of the global environment” at stake in a nuclear exchange and expressing “grave concern over the lack of progress in completing nuclear arms reduction agreements.”622

Scientists soon began to build on environmentalists’ concerns. In 1982, the

Swedish Academy of Sciences asked Paul Crutzen, then at the Max Planck Institute for

Chemistry, to look into the potential effects of a nuclear exchange on the ozone of the upper atmosphere for a special edition of the scientific journal Ambio that dealt specifically with the environmental consequences of nuclear war.623 Ozone had become a popular and important environmental issue both in Europe and the United States in the early 1980s, and Ambio’s editors recognized that a global nuclear war would inject

enormous amounts of ozone-depleting NOx into the atmosphere. Strong supporters of disarmament, the Swedish Academy hoped to use this stratospheric impact to reinforce the global stakes of a bilateral nuclear conflict between the U.S. and the Soviet Union.624

621 As the International Union for the Conservation of Nature affirmed in 1981, “peace is a contributory condition to the conservation of nature, just as conservation itself contributes to peace through the proper and ecologically sound use of natural resources.” From Memo, Sierra Club Task Force on Environmental Impacts of Military Projects to Chapter Chairs, Conservation Chairs, Group Chairs, RCC Chairs, Council Delegates, BOD, (no date, c. 1981), International Committee, Box 3, Folder 19, Sierra Club International Program Records, Sierra Club Archives, Bancroft Library, University of California, Berkeley. 622 “Sierra Club Task Force on Environmental Impacts.” 623 The volume represented the culmination of a two-year project by the Academy, supported, in part, by the Sierra Club. Ambio press conference invitiation, Siera Club International Program Files, Folder 30, Box 9, Sierra Club Archives, Bancroft Library, University of California, Berkeley. 624 Conway, Atmospheric Science at NASA, 208; Badash, “Nuclear Winter,” 80.

264 Crutzen went Ambio one better. In 1980, UC Berkeley’s Nobel Laureate physicist

Luis Alvarez and his geologist son, Walter, had published a controversial study that attributed the extinction of the dinosaurs to climatic changes resulting from ash and dust sent into the stratosphere by a 10-kilometer wide asteroid some 65 million years ago.625

Crutzen and his Ambio coauthor John Birks of the University of Colorado realized that though a number of scientists had studied the immediate thermal impacts of nuclear blasts, nobody had yet accounted for the climatic impacts of the soot, smoke, and other particulate matter that a nuclear war and the resulting fires could distribute through the various levels of the atmosphere. “The Atmosphere After a Nuclear War: Twilight at

Noon” considered the possibility that a nuclear war could have the same sort of climatic effects as the Alvarez’ asteroid 65 million years earlier. Its title dramatically printed over a picture of a mushroom cloud, Crutzen’s paper also suggested that the consequences of this climate change might again be the extinction of the Earth’s dominant species—then dinosaurs; now, humans.626

In the spring of 1982, a small group of planetary scientists at NASA’s Ames

Research Center in California made up of Brian Toon, Tom Ackerman, and Jim Pollack took up the Crutzen and Birks hypothesis and began to work it into their own atmospheric models. Supported by Carl Sagan of Cornell University and Richard Turco of defense contractor Research Associates, the Ames group had already begun to apply

625 Luis W. Alvarez, Walter Alvarez, Frank Asaro, and Helen V. Michel, “Extraterrestrial Causes for the Cretaceous-Tertiary Extinction,” Science, vol. 208, no. 4448 (Jun. 6, 1980), pp. 1095-1108. See also Conway, Atmospheric Science at NASA, 207; Luis W. Alvarez, Alvarez: Adventures of a Physicist (New York: Basic, 1987); Trevor Palmer, Perilous Planet Earth: Catastrophes and Catastrophism Through the Ages (New York: Cambridge University Press, 2003); C.G. Wohl, “Scientist as Detective: Luis Alvarez and the Pyramid Burial Chambers, the JFK Assassination, and the End of the Dinosaurs,” American Journal of Physics, vol. 75, no. 11 (November, 2007): 75. 626 Paul J. Crutzen and John W. Birks, “The Atmosphere After a Nuclear War: Twilight at Noon,” Ambio, vol 11, no. 2/3 (Winter, 1982): 114-125.

265 their knowledge of planetary atmospheres and terrestrial vulcanism to both the controversial Alvarez asteroid thesis and the climatic impacts of a nuclear exchange.627

Inspired by the Crutzen paper, the “TTAPS” group (Turco, Toon, Ackerman, Pollack, and Sagan) stepped up their research on the atmospheric impacts of a nuclear war and began to use their one-dimensional radiative-convective model, along with publicly available information on things like nuclear yields and damage estimates, to simulate the potential climatic impacts of a range of nuclear exchange scenarios.628 Designed to handle multiple scenarios relatively quickly, the simple model did not have an ocean (an important moderating force in the climate system, as the authors recognized), nor did it account for seasons or horizontal circulation.629 Nevertheless, the TTAPS results strongly suggested that even a comparatively modest nuclear exchange could, as Crutzen suggested, lead to major atmospheric changes, including a massive depletion of ozone and the sudden onset of subfreezing temperatures during the summer.630 Forbidden by

NASA to use “nuclear war” in the title of any published material, Turco labeled these cold, dark, apocalyptic conditions “nuclear winter.”631

Initially, the TTAPS group took great pains to distance their work from defense politics. They feared that too close a connection to the MX missile controversy and

627 Badash, “Nuclear Winter,” 80. 628 Badash, “Nuclear Winter,” 79-80; Conway, Atmospheric Science at NASA, 209. 629 For more on climate models, see Chapter 4. 630 As Conway explains, the Ames group actually came up with the same overall result as Crutzen—a dramatic decrease in temperature—but for atmospherically different reasons. Whereas Crutzen believed that the most important changes would occur in the stratosphere, as in the Alvarez asteroid scenario, the Ames model showed that the smoke and soot in the troposphere (the lower atmosphere) would absorb enough heat to create a global temperature inversion. Little or no heat would reach the lower atmosphere, and conditions on the ground would be cold and dark, but the stratosphere would actually warm in this scenario. Conway, Atmospheric Science at NASA, 209. 631 As Badash relates, NASA insisted there be no discussion of casualties and that “nuclear war” not appear in the title. As a civilian agency, these national defense concerns lay at the fringes of NASA’s scientific jurisdiction; more importantly, NASA administrators had no intention of drawing a powerful president’s ire with bad news on his defense policy. Badash, “Nuclear Winter,” 87.

266 Reagan’s rearmament putsch might damage their scientific credibility and jeopardize their government funding. They had cause for concern. When Pollack agreed to chair an

American Geophysical Union session on “Climatic Variations on the Terrestrial Planets” that included an abstract on nuclear winter in 1982, for example, Ames administrators forced the Ames group to withdraw the paper on “bureaucratic” grounds.632 Leaders in

NASA gave specific instructions for what the civilian agency could and could not support with money and computer time vis-à-vis defense-related research, and they made it clear that TTAPS stood very close to the edge. As Turco later recalled, “there was tremendous pressure for us to not get involved.”633

Carl Sagan, however, was nothing if not politically involved.634 A talented science popularizer and one of the more famous scientists alive in the United States at the time, Sagan hoped to capitalize on the political implications of the Ames group’s findings as a way to demonstrate the folly of rearmament. In 1982, representatives of the

Rockefeller Family Fund, the Henry P. Kendall Foundation, and the National Audubon

Society approached Sagan to help organize a conference on the long-term consequences

632 Technically, Ames’ deputy director Angelo Gustafero had a point; the paper had not cleared an internal NASA review, a necessary step on the road to publication. Presenting papers at conferences like the AGU, however, typically went forward without such permission, and Gustafero himself admitted that he squelched the paper in part because he feared a budgetary response from the Reagan Administration. Later, however, when the Reagan Administration finally did cut funding for NASA’s nuclear winter research in the mid-1980s, Ames administrators kept the program alive by allocating Toon, Pollack, and Ackerman available discretionary funding and computer time. Badash, “Nuclear Winter,” 85; R. Jeffrey Smith, “‘Nuclear Winter’ Feels Budgetary Chill,” Science, vol. 227, no. 4689 (February 22, 1985): 294-5. 633 Quoted in Badash, “Nuclear Winter,” 86. 634 In Science as a Contact Sport, Stephen Schneider remembers that Sagan’s fame had not always been associated with political positions; quite the contrary, Schneider remembers urging Sagan to use his power as a public figure to press for the important political causes he believed in. Nuclear winter was the first of those major causes, and though Schneider takes some credit implicitly for urging Sagan to take up arms, both Schneider and Sagan’s biographers Ray Spangenburg and Kit Moser argued that Sagan’s liberal and politically active new wife, Ann Druyan, pushed him into the political fray more than anyone else. See Schneider, Contact Sport, 97-98; Ray Spangenburg and Kit Moser, Carl Sagan: A Biography (Westport, CT: Greenwood Press, 2004. See also Tom Head (ed), Conversations with Carl Sagan (Jackson, MS: University Press of Mississippi, 2006); Keay Davidson, Carl Sagan: A Life (New York: Wiley, 2000); Richard P. Turco, “Carl Sagan and Nuclear Winter,” in Carl Sagan’s Universe, edited by Yervant Terzian and Elizabeth Bilson (New York: Cambridge University Press, 1997): 239-46.

267 of nuclear war. Sagan established a steering committee comprised of Stanford biologist

Paul Ehrlich, NCAR founder Walter Orr Roberts, and Brookhaven’s George Woodwell, and the committee soon secured the financial support and political backing of a wide variety of scientific and environmental groups, including NRDC, the Environmental

Defense Fund, and the Sierra Club. Led by Sagan, the steering committee made the

TTAPS nuclear winter paper, then still in progress, the centerpiece of their efforts.635

Sagan’s group began by working to establish a wide-spread agreement on the basic phenomenon of nuclear winter within the scientific community. In April of 1983, with the draft of what would become known as “the TTAPS paper” all but complete,

Sagan invited a group of a hundred prominent scientists from a variety of fields to assess the group’s model and its conclusions in a private workshop on the issue in Cambridge,

Massachusetts.636 The conference had three purposes. First, Sagan legitimately hoped to flesh out any fatal flaws in the theory or the data that might undermine the nuclear winter hypotheses before he introduced the controversial idea to a broader public. Second, by asking other scientists to comment on the work and incorporating their suggestions into the TTAPS results, Sagan hoped to establish a buy-in from a community of prominent scientists who could back him if critics tried to undermine his policy position by challenging his science. Finally, the Cambridge workshop served as a challenge for other groups working on the scientific problem of nuclear winter—Stephen Schneider, Starley

Thomson, and Curt Covey at NCAR and a group led by Mike MacCracken and Cecil

“Chuck” Leith at Lawrence Livermore National Laboratory in particular—to move forward with their own research, which Sagan and his colleagues believed would only

635 Conway, Atmospheric Science at NASA, 210. 636 Conway, Atmospheric Science at NASA, 210; Badash,”Nuclear Winter,” 80.

268 strengthen the TTAPS study.637 Despite some discrepancies between model results and a good deal of uncertainty about the basic assumptions of both nuclear defense strategy and atmospheric processes among the participants, Sagan saw this private review and the research that came out of it as a vote of confidence for the TTAPS hypothesis—and for the subsequent public conference.

With atmospheric scientists presumably on board after Cambridge, Sagan and his colleagues orchestrated the Washington, D.C. “Conference on the Long-Term Worldwide

Biological Consequences of Nuclear War” to garner as much public and political attention as possible.638 The steering committee scheduled the two-day event to begin on

Halloween, 1983. On October 30, Sagan published an exposé on nuclear winter in

Parade Magazine, a popular Sunday newspaper supplement with more than 20 million readers. Chaired by George Woodwell and kicked off by Stanford University’s eloquent

President, Donald Kennedy, the conference itself ran less like a scientific meeting than like an extended, staged scientific press release.639 A satellite link—relatively new technology in 1983—connected an audience of several hundred scientists, press, and politicians to members of the Soviet Academy of Sciences in Moscow.640 Sagan, drawing

637 George Woodwell to Stephen Schneider, July 11, 1983, Nuclear Winter Files, Folder 2, Box 27-1 (3 of 4), Stephen Schneider Papers, UCAR/NCAR Archives, Boulder, CO. 638 See Paul Ehrlich, Carl Sagan, Donald Kennedy, and Walter Orr Roberts, The Cold and the Dark: The World After Nuclear War, Conference on the Long-Term Worldwide Consequences of Nuclear War, Washington, D.C., 1983 (New York: Norton, 1984); Paul R. Ehrlich et al., “Long-Term Biological Consequences of Nuclear War,” Science vol. 222, no. 4630 (December 23, 1983): 1293-1300. 639 Donald Kennedy, “Conference on the Long-Term Biological Consequences of Nuclear War: Introduction,” draft, October 24, 1983, Nuclear Winter Files, Box 5033 (1 of 4), UCAR/NCAR Archives, Boulder, CO. In The Discovery of Global Warming, Weart actually refers to the event as a press conference rather than a scientific meeting; in truth it was some combination of the two. Weart, Discovery, 143-44; Badash, “Nuclear Winter,” 79. 640 As Tom Malone’s draft remarks reveal, the satellite link was relatively well scripted, except that the Americans didn’t know exactly who they’d be talking to. Malone’s October 26 draft included not only his own opening comments, but also parenthetical instructions for camera direction and the expected physical responses of his Soviet counterparts…whomever they may be. Malone, “The Moscow Link: Opening

269 on a career in popular science, delivered a concise, accessible, and alarming synthesis of the TTAPS work, which he reinforced with a well-produced video, The World of Nuclear

Winter.641 Ehrlich, also a well known scientist, not only for The Population Bomb but also for his frequent appearances on Johnny Carson’s Tonight Show, then gave a summary of a separate study on the potential biological impacts of a nuclear exchange and a subsequent nuclear winter that he and his biologist colleagues had conducted in parallel with the TTAPS research.642 On November 1, Sagan and Ehrlich followed up with an appearance on Ted Koppel’s “ABC News Nightline.”643 As intended, their

“scientific congress” made headlines around the world.644

In Washington, the conference organizers insisted that they had “rigorously avoided drawing any policy implications from their findings,” but Sagan had a detailed and specific policy response in mind, which he published in the journal Foreign Affairs shortly after the TTAPS study appeared in Science in December of 1983.645 The TTAPS

Remarks,” Draft, October 26, 1983, Nuclear Winter Files, Box 5033 (1 of 4), UCAR/NCAR Archives, Boulder, CO. 641 Schneider, Contact Sport, 102. 642 See Ehrlich et al., The Cold and the Dark. 643 Badash, “Nuclear Winter,” 89. 644 For example, Philip Shabecoff, “Grimmer View is Given of Nuclear War Effects,” The New York Times, October 31, 1983; Terry Atlas, “A Grim View of Nuclear War; Scientists Say it would Leave Earth Cold, Dark,” The Chicago Tribune, October 31, 1983; “Scientists See Man’s Doom in Nuclear Winter,” The Los Angeles Times, December 8, 1983; “Biologists Paint Icy Picture of How the World Could End,” The Washington Post, November 1, 1983. 645 For some of the scientists who had helped produce the studies, the statement of neutrality probably rang true. For Kennedy, Sagan, Ehrlich, Roberts, and Woodwell, however, that was clearly not the case, even at the Washington, D.C. conference. In his opening remarks, Kennedy put the problem in stark historical terms and spelled out, albeit vaguely, what the group hoped for in a policy response. Kennedy called on the President, Congress, the Soviet Union, and other foreign leaders (and, curiously, the NAS) to: 1) “intensify substantially” efforts to achieve new international agreements on nuclear weapons; 2) to “take all practical actions that could reduce the risk of nuclear war by accident or miscalculation;” to take measures to “inhibit the further proliferation of nuclear weapons to additional countries;” 3) to uphold existing arms control agreements, including SALT II; and 4) “to avoid military doctrines that treat nuclear explosives as ordinary weapons of war.” Kennedy, “Introduction,” 10/24/83; the quote in the main text is from Malone, “The Moscow Link,” 10/26/83. Carl Sagan, “Nuclear War and Climate Catastrophe: Some Policy Implications,” Foreign Affairs, vol. 62, no. 2 (Winter, 1983-84): 257-92; R. Turco, O. B. Toon, T. Ackerman, J. B. Pollack, and C. Sagan, “Nuclear Winter: Global Consequences of Multiple Nuclear

270 one-dimensional model identified a “crude threshold, around 500 to 2,000 warheads” for a nuclear exchange that could trigger a nuclear winter.646 Sagan used this threshold as a way to examine the potential of existing postures on nuclear defense and disarmament for dealing with the prospect of nuclear winter. He addressed eight different defense strategies, from new treaties on targeting and yields supported by moderates and liberals to Reagan’s proposed ballistic missile defense system (the centerpiece of SDI, or, pejoratively, “star wars”). Not surprisingly, given that an exchange of as few as 10% of the existing nuclear arsenals of the U.S. and Soviet Union could trigger a climatic doomsday scenario, Sagan found all of these prospects lacking. “None of the foregoing possible strategic and policy responses to the prospect of a nuclear war-triggered climatic catastrophe seem adequate even for the security of the nuclear powers,” he wrote, “much less for the rest of the world.”647 Moreover, he contended, in a world now linked by the potential of climatic disaster, “beyond the climatic threshold, an increase in the number of strategic weapons leads to a pronounced decline in national (and global) security.”648

With the long-term effects of even a limited nuclear war—a dubious concept for

Sagan—promising mutual suicide, Sagan argued that only a build-down of nuclear stockpiles to levels below the nuclear winter threshold could restore a logically sound and politically credible policy of deterrence.649 “For me,” he wrote, “it seems that the species is in grave danger at least until the world arsenals are reduced below the threshold for climatic catastrophe.”650

Explosions,” vol. 222, no. 4630 (December 23, 1983): 285. 646 Sagan, “Nuclear War and Climate Catastrophe,” pg. 286. 647 Ibid., 283 648 Emphasis in original. Ibid., 286 649 Ibid., 284. 650 Ibid.

271 Despite its rhetorical punch, Sagan’s Foreign Affairs article had little impact on the politics of disarmament. His ongoing efforts to parlay the science of nuclear winter into defense policy had a profound effect on the politics of climate science, however. Not surprisingly, both the style and substance of Sagan’s arguments about nuclear winter irked conservative scientists. Many objected to the unorthodox presentation of scientific materials in Parade and Foreign Affairs, and they balked at Sagan’s close association with Ehrlich, an outspoken liberal and environmentalist. Long-time Republican science administrators like Robert Jastrow and Fred Singer, alongside one time NAS President

Russell Seitz, feared that activists like Sagan and Stephen Schneider might push the field too far into the wrong kind of politics. When Science ran an editorial in front of the

TTAPS paper praising Sagan’s and Ehrlich’s work on the issue in December of 1983, conservatives both within the scientific community and outside of it were incensed at what they saw as a left-wing agenda at the journal. In response, Jastrow and a small group of colleagues, primarily scientists associated with the defense industry, formed the

George C. Marshall Institute, a conservative think tank initially aimed at supporting SDI

651 and nuclear energy that soon became vocal on both the ozone and CO2 issues.

Meanwhile, Singer and Seitz both publicly attacked Sagan’s policy recommendations not only by challenging the TTAPS group’s scientific results, but also by impugning their methodology and openly questioning their motives.652

651 Conway, Atmospheric Science at NASA, 212. See also Oreskes et al., “Chicken Little”; Naomi Oreskes and Erik M. Conway, “Challenging Knowledge: How Climate Change Became a Victim of the Cold War,” in Agnotology: The Making and Unmaking of Ignorance, edited by Robert N. Proctor and Londa Schiebinger, (Stanford: Stanford University Press, 2008): 55-89. 652 Seitz questioned the qualifications of the scientists who had confirmed the TTAPS study at the Cambridge conference; Singer accused Sagan of using unproven assumptions and cherry-picking results from worst-case scenarios. They both couched their criticism as scientific, but ultimately they couldn’t help but also attack Sagan’s policy approach. S. Fred Singer, “The Big Chill? Challenging a Nuclear

272 Sagan anticipated these objections, and for the most part he and his colleagues responded to them professionally and convincingly.653 He did not anticipate the criticism he soon received from scientists who generally shared his political beliefs, however. His highly publicized disagreement with Stephen Schneider in particular further fed conservative’s claims that Sagan had exaggerated the TTAPS results.

Schneider began studying the nuclear winter issue with Curt Covey and Starley

Thompson at NCAR in the spring of 1983 with funding from the Defense Nuclear

Agency. In loose collaboration with the TTAPS group and the Lawrence Livermore

Laboratory, the NCAR team sought to create a more realistic picture of the climatic consequences of a nuclear exchange by recreating the TTAPS scenarios on a three dimensional climate model. Whereas the TTAPS study derived their conclusions from a model of a vertical column of “dead air,” Schneider and his group worked to incorporate seasonal variations, horizontal mixing, and the stabilizing force of the oceans into their model.

Almost immediately, NCAR’s results began to differ from those of the TTAPS study. Qualitatively, the NCAR model supported the phenomenon: in the event of a nuclear exchange, in all likelihood smoke, soot, and dust would lead to a catastrophic cooling of large portions of the Earth. But the details of the three dimensional model revealed a more complex and somewhat less stark range of specific climatic responses.

Scenario,” The Wall Street Journal, February 3, 1984; Russell Seitz, “Comments and Correspondences,” vol. 63, no. 4 (Spring, 1984): 998-999. 653 Singer essentially reprinted his Wall Street Journal editorial—or at the very least rehashed its arguments—in Nature and Science in the ensuing two years, and Sagan, along with a group of colleagues including Steve Schneider and Starley Thompson, among others, continued to rebut his attacks. See S.F. Singer, “Is the ‘Nuclear Winter’ Real?” Nature 310 (August 16, 1984): 625; Edward Teller, “Widespread After-Effects of Nuclear War,” Nature 310 (August 16, 1984): 621-624; S. Fred Singer, Cresson H. Kearny, R. P. Turco, O. B. Toon, T. P. Ackerman, J. B. Pollack and C. Sagan, “On a ‘Nuclear Winter,’” Science, vol. 227, no. 4685 (Jan. 25, 1985), pp. 356-444.

273 A nuclear war in the summertime had nearly the same severity of consequences in both the one dimensional and three dimensional models, but a January strike in NCAR’s 3-D model (a scenario impossible to simulate in TTAPS model) yielded a much more minor set of impacts. In all of the 3-D scenarios, the oceans significantly modulated the overall temperature change, especially near the coasts, and ultimately the climatic result that

Schneider and his colleagues arrived at resembled the variability of a severe and unpredictable autumn more than a deep winter freeze. Perhaps most importantly, the variability revealed by Schneider’s 3-D model—the “real world,” as he called it—seemed to undermine Sagan’s “crude threshold” for perpetual winter.654

Between the fall of 1983 and the summer of 1986, Sagan and Schneider engaged in a series of exchanges on the issue. Initially, Schneider privately tried to convince Sagan to back off of the threshold idea that he, Schneider, believed to be an artifact of the 1-D model. Schneider shared Sagan’s general bent toward disarmament, and like Sagan, he hoped that the climatic effects of a nuclear exchange would help demonstrate the insanity of the “winnable” nuclear war concept. An outspoken liberal who had encouraged Sagan to use his popular appeal to take more of a political stand on scientific issues as early as

1977, Schneider certainly believed in using science as a political tool.655 Moreover, he and Sagan jointly opposed the Administration’s response to the issue from the beginning.

In early 1985, Reagan’s people attempted to defuse the issue by co-opting it. In a dismissive report, the Department of Defense claimed to have incorporated nuclear winter into existing defense strategy; Weinberger went so far as to suggest that nuclear

654 Schneider, Contact Sport, 100-101; Conway, Atmospheric Science at NASA, 211. 655 Schneider, Contact Sport, 97.

274 winter actually strengthened the case for SDI.656 Throughout 1984 and 1985, Schneider and Thompson spent a good deal of time and ink defending the nuclear winter hypothesis and reiterating its potential policy impact, despite their problems with some of Sagan’s science.

But for Schneider, politically motivated science had to be impeccable, and popular scientists like Sagan had a responsibility to alter the specifics of their political objectives as the science behind those objectives changed. When Sagan refused to budge on the threshold concept, Schneider felt he had to air his results for the sake of scientific credibility. Over the next two years, the TTAPS group and Schneider’s NCAR group engaged in a series of scientific exchanges that ranged from scientific journals like

Science and Nature to the popular press and, again, the pages of Foreign Affairs.

Schneider and Thompson argued that “the global apocalyptic conclusions of the nuclear winter hypothesis can now be relegated to a vanishingly low level of probability.”657

Nuclear winter thus could not provided the sole impetus for a build-down policy as Sagan had described. Still, as Schneider pointed out, even nuclear fall could have a catastrophic effect on world agriculture, and he continued to urge the defense community to incorporate the environmental consequences of nuclear war more meaningfully into

656 Weinberger’s Assistant Secretary of Defense, Richard Perle, spelled out the unimportance of nuclear winter to Administration defense policy in hearings before the Senate Armed Services Committee in October of 1985. Caspar W.Weinberger, “The Potential Effects of Nuclear War on the Climate: A Report to the United States Congress,” March, 1985, in U.S. Congress, Senate Armed Services Committee, Nuclear Winter and its Implications, Hearing, 99th Congress, 1st Session, October 2 and 3, 1985 (Washington, D.C.: U.S. Government Printing Office, 1985): 72-90. See pg. iii of the report for Weinberger’s response; 133-4 of the hearing for Perle’s remarks. 657 One of Sagan’s compelling arguments had to do with the possibility of human extinction, which Schneider and Thompson had been skeptical about since as early as the spring of 1983. Using the “average” time of existence for a “successful” species, Sagan estimated that nuclear weapons, through the generations they prevented, would actually cost somewhere along the lines of 500 trillion lives over time. Sagan, “Nuclear War and Climatic Catastrophe,” pg. 275; Starley L. Thompson and Stephen H. Schneider, “Nuclear Winter Reappraised,” Foreign Affairs vol. 64, no. 5 (Summer, 1986): 981-1005. For the quote, see pg. 983.

275 strategic planning. For Schneider, like Sagan, this included “significantly reduced levels of arsenals.”658

As with the 1983 EPA and NAS reports on CO2, the disagreement between

Schneider and Sagan on nuclear winter allowed skeptics and political opponents to cast the whole issue as speculative. Schneider repeatedly insisted that he and Sagan had “few fundamental differences,” but conservative climate scientists, skeptical defense analysts, and the press used the areas where the NCAR and TTAPS models disagreed to highlight the uncertainties of the hypothesis and to undermine the credibility of the science.659

Conservative commentators echoed conservative scientists in accusing the greater climate science community of taking too cavalier an approach to their evidence.660 They criticized liberals like Sagan and Schneider for going public with their results before they had fully fleshed out the science. As Edward Teller, a nuclear winter skeptic and a the primary booster of SDI, commented in Nature, “highly speculative theories of worldwide destruction—even of the end of life on Earth—used as a call for a particular kind of political action serve neither the good reputation of science nor dispassionate political thought.”661

Though the hypothesis itself had little to do with CO2, the nuclear winter debate served as a point of convergence for the increasingly intertwined science and politics of climate change. From the beginning, liberal scientists, environmentalists, and politicians gravitated toward nuclear winter because of its political implications. Much in the same

658 Thompson and Schneider, “Nuclear Winter Reappraised,” 984. 659 See U.S. Congress, Nuclear Winter, Hearing, 1985. 660 Schneider’s falling out with Sagan may have reflected his experiences with similar accusation that met the publication of his first book, The Genesis Strategy, in the late 1970s. In any case, as he recounts in Contact Sport, in 1983 Schneider did his best to convince Sagan not to go public until there was broader agreement on the threshold issue. He remembers a “pact” the scientists made—a pact that Sagan broke in Washington on Halloween. Schneider, Contact Sport, 100. 661 Edward Teller, “Widespread Aftereffects,” pg. 624.

276 way that they had mobilized CO2 research to undermine the Administration’s energy policy, these same groups—constituted largely of the same individuals—studied and publicized the climatic impacts of a nuclear war as a form of public resistance to

Reagan’s defense policy, and to the Administration more generally. Again, this dissent played out in congressional hearings (again sponsored by Al Gore), at scientific conferences and in scientific journals, and within the bureaucracy. To a much greater

extent than the CO2 issue did in the early 1980s, nuclear winter also appeared in public media like newspapers, television, and magazines. And finally, again, the experience drove liberal climate scientists, Democratic politicians, and leaders of environmental organizations to cooperate in response to a common opponent: the Reagan

Administration. The organizers of the Halloween conference worked in offices in

Washington provided by Gus Speth and his new World Resources Institute. Speth’s former employer, NRDC, meanwhile worked to obtain documents on the U.S. Navy and

Defense Nuclear Agency’s response to nuclear winter—documents the organization distributed to both liberal scientists like Schneider and to Congress.662

Studying the climatic impacts of nuclear war provided climate scientists with a

more subtle way of resisting the Administration’s position on CO2-induced climate change as well. The nuclear winter debate gave many of the same scientists who saw their funding from the DoE cut under Reagan new access to government money through the Department of Defense.663 The DoD-sponsored research paid scientific dividends,

662 Badash, “Nuclear Winter,” 87. Department of the Navy to Estelle H. Rogers (NRDC), March 14, 1984, and J.A. Lyons, “Memorandum to the Chief of Naval Operations: The World After a Nuclear War,” November 7, 1983, in Nuclear Winter Files (Box 3 of 4), Stephen Schneider Papers, UCAR/NCAR Archives, Boulder, CO. 663 MacCracken’s climatic research at Lawrence Livermore, for example, had originally received funding primarily from the DoE, which essentially ran non-classified portion of the lab. The nuclear winter

277 especially for climate modelers who used nuclear winter as a test case—a “digital thought experiment,” as Erik Conway calls it—for their rapidly developing models of atmospheric radiation and circulation.664 The ongoing controversy fostered a new level of inter-institutional collaboration that helped in the development of the Community

Climate Model (so named for its nearly universal usage), and scientists’ experiences working with the model helped contribute to new questions not just about global circulation, climate sensitivity, and climate forcing, but also about the specific impacts of these phenomena—an area specifically targeted in the DoE cuts.665 The new DoD money thus helped climate scientists circumvent Reagan’s funding cuts for climate research—research that ultimately provided the raw material for more activist scientists’ attacks on the very Administration that had cut the DoE research budget in the first place.

Nuclear winter research was in this sense not just form of dissent, but also a subtle form of retaliation.

Conclusion

Alongside the ongoing debate over CO2, the nuclear winter saga at once reflected and helped to shape an increasingly complex relationship between the science and politics of climate change in the 1980s. Since the beginning of the Cold War, American scientists had become accustomed to three loose and often overlapping categories of political activism, each controversial in their own right. First, the majority of scientists’ research gave him access to DNA and DoD money, and essentially saved his climate program. Schneider’s group, too, benefited from DNA money. Stephen Schneider to Michael MacCracken and Cecil Lief, December 23, 1983, DoD/LLNL Proposal Folder; Memo for Stephen Schneider from the Office of the Secretary of Defense, May 29, 1984, Defense Science Board Task Force Folder; Bill Hess, Memo to Steve Schneider and Bob Dickinson, December 19, 1983, DNA Proposal Folder, all in Box 5033 (1 of 4), Nuclear Winter Files, UCAR/NCAR Archives, Boulder, CO. 664 Conway, Atmospheric Science at NASA, 211. 665 Schneider to MacCracken and Leith, 12/23/83; Memo for Schneider, 4/29/94; Hess Memo 12/19/83.

278 activism revolved around efforts to influence science policy—and in particular, government policies on science funding. Beginning in the 1950s, scientists worked primarily with government money, and they had to continually convince administrators at

OSTP, the NSF, and various government agencies that their research served the government’s interests, and, ultimately, the public good. Tensions arose over the proper balance between these political realities and the scientific ideals of political neutrality and independence, and these tensions played out differently within and between different disciplines and communities in American science.666 Second, scientists often demonstrated the value of their research by framing their work as particularly applicable to specific practical decisions and policy discussions outside of science. Scientists frequently served as expert advisors in legal and policy debates without necessarily overtly taking sides. Finally, as the American physics community demonstrated repeatedly in their advocacy on nuclear defense issues throughout the Cold War, many scientists used their professional status to gain credibility as advisors and advocates on policy issues either unrelated or only tangentially related to their expertise.667 Here they traditionally cast themselves not as experts per se, but rather as concerned citizens with a particularly valuable intellectual perspective.

As scientists aligned with Congressional Democrats and environmentalists in opposition to Reagan’s defense and energy policies, these three forms of advocacy bled

666 See chapter 4. 667 Badash, “Nuclear Winter,” 103. See also Badash, A Nuclear Winter’s Tale. For more on the physics community’s role in defense politics, see Badash, Science and the Development of Nuclear Weapons: From Fission to the Limited Test Bad Treaty, 1939-1963 (Atlantic Highlands, NJ: Humanities Press, 1995); Daniel Kevles, The Physicists: The History of a Scientific Community in America (Cambridge, MA: Harvard University Press, 1995); Gregg Herkin, Brotherhood of the Bomb: The Tangled Lives and Loyalties of Robert Oppenheimer, Ernest Lawrence, and Edward Teller (New York: Henry Holt, 2002); McGeorge Bundy, Danger and Survival: Choices About the Bomb in the First Fifty Years (New York: Random House, 1988); Zuoyue Wang, In Sputnik’s Shadow: The President’s Science Advisory Committee and Cold War America (New Brunswick, NJ: Rutgers University Press, 2008).

279 together until they could scarcely be distinguished. In the debate over CO2-induced warming, scientists went out of their way to convert small changes in science itself into bold arguments for changes in policy. The tactic both strengthened the case for further research funding and buttressed existing arguments for rethinking the nation’s priorities in developing new energy strategies. Climate scientists’ role in nuclear winter further challenged the boundaries of science activism. Sagan, for example, not only interpreted scientific results in terms of existing policy options; he used his own research as the rationale for a new and more aggressive policy of disarmament—an area certainly outside of his scientific training. Singer’s editorial responses, in turn, attacked Sagan from both a

668 scientific and a military perspective. In the case of both nuclear winter and CO2, the new political salience of climate research made the prosecution and presentation of climate science itself a form of political act. For Hansen, Schneider, and Sagan, it was an act of political defiance against an unfriendly Administration; for Singer, Seitz, and perhaps to a lesser extent, Nierenberg, it was an act of political solidarity with that

Administration.669

It is perhaps unfair to paint the climate science community’s political commitments in the 1980s with such a broad brush; indeed, to scientists concerned about the credibility of the profession, generalizations based on the specific activities of outspoken individuals may seem invidious. As Schneider’s account of his own personal struggle with the tension between his political beliefs and his professional responsibility bears testament to,

668 Singer, The Wall Street Journal, 2/3/84; Singer, “Is the ‘Nuclear Winter’ Real?” 669 One interesting difference between the two groups was their role in the scientific process. Liberal scientists like Schneider, Hansen, and Revelle typically not only promoted climate science, they also created some of the most cutting edge work on the subject throughout their careers. With the partial exception of Nierenberg, conservative scientists from the Marshall Institute and elsewhere had won scientific credibility in related fields, but spent much more ink tearing down novel climate science and managing or challenging the consensus view than they did producing their own original work on the subject. See Oreskes and Conway, “Chicken Little” and Merchants of Doubt.

280 a deep ambivalence about the relationships between science and policy among the discipline’s rank and file accompanied its polarization in the 1980s. That ambivalence persists in memory. Every scientist’s experience represented a different mix of personal and professional values, and few remember the episode without expressing some degree of emotion. The battle to incorporate the results of climatic research into energy and defense policy certainly divided the community, but the contours of those divisions were neither static nor clean. Ultimately, however, within the variety of intense individual

responses to nuclear winter and the CO2 debate lie a common recognition that between

1979 and 1985 the politics of atmospheric science had fundamentally changed.

The tone of the climate debate took such a combative turn in the 1980s in part because both scientists and politicians began to recognize not only the social and environmental consequences of global warming, but also the potential political consequences of the issue, especially as it applied to energy. Still, while scientists,

environmentalists, and democrats used the environmental impacts of CO2 as a way to criticize and embarrass the Reagan Administration for its irresponsible positions on science research and renewables, no group offered any specific policy solutions to the problem of global warming itself, at least not beyond the conservation efforts they were already pushing. Despite the international environmental community’s best efforts to create a framework for handling global environmental problems, no legal or regulatory mechanisms existed for dealing with the global causes or consequences of climate change. The domestic political resistance to even the suggestion that solutions could involve changes in the fossil fuel energy mix did not bode well for more detailed responses in the future, either domestically or abroad. Especially after the Global 2000

281 report, groups like NRDC, the World Resources Institute, the Sierra Club, and Friends of the Earth began to recognize that alongside acid rain and ozone, global warming represented both an ecological and a political game changer. With an unfriendly

Administration at home, these groups looked toward scientists and the international political community to help define how exactly this new game would be played.

282 Chapter 6 Mechanisms of Change: Knowledge and Regulation in a Warming World

It took American environmentalists until the middle of the 1980s to commit significant financial, human, and political resources to the issue of global warming.

When they did, they found themselves dealing with exactly the type of practical problems that Michael McCloskey warned against in his 1982 rubric for international environmental campaigns.670 The global scale of climate change and the protracted chronology of its potential impacts transcended the local, regional, and national political and legal frameworks that environmentalists relied on in order to affect change on most other environmental issues.671 Global warming affected discrete environmental spaces in complex ways and only through long chains of causation. Because of its borderlessness and the highly technical nature of its problems, the global atmosphere had no obvious constituency outside of science. Scientists continued to couch their conclusions on global warming in the caveats of scientific uncertainty, and opponents of energy and fossil fuel regulation used these uncertainties to dismiss the problem. Despite Al Gore’s hearings on the issue in the early 1980s, by 1985 only a small fraction of the American public

recognized CO2 as an environmental problem, and many conflated CO2-induced warming with other atmospheric problems like ozone depletion and acid rain. Both scientists and environmentalists recognized a need for action on climate change in the 1980s, and popular support for such an effort did begin to grow over the course of the decade. But before they could fight global warming, scientists, environmentalists, and their political

670 Michael McCloskey, “Criteria for International Campaigns,” December 1, 1982. International Committee, 1972-1983, Meetings and Conferences, Operational Files, Sierra Club International Program Records [3:19]. See Chapter 4. 671 Gerald O. Barney, Global 2000 Report to the President: Entering the 21st Century (Washington, D.C.: U.S. Government Printing Office, 1980).

283 allies had to develop a new set of global political strategies and international regulatory tools that could accommodate the scale and complexity of the problem.

Advocates of action on global warming built their international political response on the model of remarkably successful efforts to tackle two related problems of the global atmosphere in the 1980s: acid rain and ozone depletion. Like global warming, both ozone depletion and acid rain challenged the legal and political boundaries of intra- national and bilateral international strategies for regulating environmental pollutants.

Working with leaders at the United Nations Environment Program and with environmental and scientific NGOs, scientists relatively quickly established an international consensus on the basic facts of each of these issues. That consensus in turn served not only as a statement of scientific fact, but also as a set of guidelines for effective international treaties regulating emissions of the gases identified as

causes—SO2 and NOx for acid rain and chlorofluorocarbons (CFCs) for ozone depletion.

Because of the global nature of ozone, the success of the 1985 Vienna Convention for the

Protection of the Ozone Layer and the 1987 Montreal Protocol on Ozone Depleting

Substances in particular gave the international scientific and political communities

confidence that they could work out a similar solution for CO2 and other greenhouse gases (GHGs).

The ozone model was problematic for climate change, however. The diffuse

nature of the sources of CO2 and the diversity and uncertainty of its potential impacts made it more difficult to establish the type of broad national-level political support for regulations on climate change that had helped to propel the international convention on ozone. Perhaps more importantly, the ad hoc process of translating scientific consensus

284 into international political action on ozone didn’t account for the high economic and

political stakes of an effort to regulate a gas—anthropogenic CO2—that came primarily from the indispensable driving force of the world economy: fossil fuel energy. Because of the potential economic impacts of regulations on energy use, efforts to create a convention on climate change drew opposition from national and corporate political actors who wielded much more power than the few chemical companies opposed to regulations on CFCs. Moreover, the very existence of the Montreal Protocol itself changed the political landscape in which a climate convention would be negotiated. If the Vienna-Montreal process provided a model for scientists and environmentalists to build a successful international legal framework based on consensus science, so too did it give politically powerful opponents of a strong international convention on climate change a better idea of what kind of process they were dealing with.

Scientific consensus provided the lynchpin for political action on ozone, and not surprisingly, environmentalists and international political leaders placed a similar premium on scientific knowledge in the development of a convention on climate change.

Here, too, global warming advocates faced the limitations of their previous successes, both on climate change itself and on related issues like ozone depletion and acid rain. At home, fierce debates over Reagan’s energy and defense policies in the 1980s had both politicized and polarized climate change discourse. Scientists, wary of attacks on their credibility at home, turned to the international scientific and political communities to help establish an unimpeachable, universal set of facts on the issue.672 They focused not only on basic climate science, but also on the potential social, economic, and environmental

672 See Chapter 5.

285 impacts of global warming, and, to a limited extent, on the pros and cons of a variety of international policy responses to the problem.673 Even working with a strong international scientific consensus, however, global warming advocates found that they lacked the political power to translate climate science into climate policy at the international scale without backing from powerful national governments. In order to garner the type of political support they needed to implement policy on climate change, they had to incorporate governments into the consensus-making process itself. To that end, scientific and environmental NGOs worked with international agencies to develop an ongoing process of politically negotiated international consensus-making on climate change eventually formalized under United Nations oversight as the Intergovernmental

Panel on Climate Change, or IPCC.

The IPCC was not officially a mechanism for making policy, however; rather, it was a mechanism for making knowledge. More specifically, scientists intended the IPCC as a way provide a certain type of knowledge—an international scientific consensus on climate change—that would serve as a guide for negotiations on a new United Nations

Framework Convention on Climate Change (UNFCC), to be introduced at the 1992

United Nations Conference on Environment and Development in Rio de Janeiro, Brazil.

Perhaps not surprisingly, the overt connection between the IPCC and the UNFCCC proved problematic. As scientists, environmentalists, and officials from the world’s governments all recognized, the IPCC first assessment report would set the terms for a set

673 World Climate Program, Report of the International Conference on the Assessment of the Role of Carbon Dioxide and of Other Greenhouse Gases on Climate Variations and Associated Impacts, WMO No. 661 (Geneva, Switzerland: World Meteorological Organization, 1986); Jill Jaeger, Developing Policies for Responding to Climate Change: A Summary of the Discussions and Recommendations of the Workshops Held in Villach (28 September-2 October 1987) and Bellagio (9-13 November 1987) under the Auspices of the Beijer Insitutie, Stockholm (Geneva, Switzerland: World Meteorological Organization; Nairobi, Kenya, United Nations Environment Program, 1988).

286 of future international negotiations dealing with high stakes issues like energy and land use that stood at the heart of the world economy. The consensus-making process offered representatives from national governments and from industry a chance to pre-empt negotiations on the UNFCCC by arguing for specific language in the IPCC report that protected their constituents’ political and economic interests. Consequently, in the run- up to the 1992 Rio “Earth Summit” and the introduction of the UNFCCC, the IPCC scientific assessment process served not only as a forum for making consensus on climate change, but also as the primary battleground for the international politics of climate change.

The Vienna-Montreal Process: Acid Rain and Ozone as Templates for Change

American environmentalists joined in the effort to regulate greenhouse gases in the 1980s in part as a response to their success in collaborating with international scientific organizations on two other global atmospheric issues. International agreements on acid rain and ozone depletion helped to convince many environmentalists that they could effectively protect the global atmosphere through new international legal and regulatory regimes under the auspices of the United Nations. These agreements—and particularly the 1985 Vienna Convention and the 1987 Montreal Protocol—both increased awareness about global atmospheric issues more generally and provided a

template for a future international legal regime for regulating CO2 and other GHGs in order to combat climate change.

For American environmentalists, acid rain presented a much more familiar set of problems than did either ozone depletion or global warming, and though newly-pressing

287 at a global scale, it was not a new issue. In an 1872 book called Air and Rain: The

Beginnings of a Chemical Climatology, the English chemist Angus Smith used the term to describe the influence of airborne matter from coal combustion and the decomposition of organic matter on the chemistry of rainwater throughout the English, Scottish, and

German countrysides.674 In the 1950s, limnologists, agricultural scientists, and atmospheric chemists took up the issue as part of independent efforts to expand basic research within each discipline, and by the late 1960s the Swede Svante Oden had begun to incorporate these studies in a developing hypothesis that linked the causes of acid rain—namely, industrial emissions of sulfur dioxide, oxides of nitrogen, and other chemicals—with the phenomenon’s environmental and public health impacts.675 Focused

primarily on Scandinavia, Oden demonstrated that European SO2 emissions adversely affected Scandinavian plant growth, freshwater fish populations, and overall water quality in the region, and his work provided part of the impetus for Sweden’s leadership in the United Nations Conference on the Human Environment in 1972.676

Oden’s studies of acid rain helped to illuminate the potential geographical disparity

between the original sources and ultimate impacts of pollutants like SO2 and NOx. Like global warming, the issue required some new ideas about transboundary environmental governance, both in Europe and in North America. Domestically, the Nixon

674 Ellis B. Cowling, “Acid Precipitation in Historical Perspective,” Environmental Science and Technology, vol. 16, no. 2 (February, 1982), 111A. 675 Ibid, 113A. See also Samuel P. Hays, Beauty, Health, and Permanence: Environmental Politics in the United States, 1955-1985 (Cambridge: Cambridge University Press, 1987), 352; Robert H. Boyle and R. Alexander Boyle, Acid Rain (New York: Schocken Books, 1983); Carolyn Curtis (ed), Before the Rainbow: What we Know About Acid Rain (Washington: Edison Electric Institute, 1980); Ernest Y. Yamarella and Randal H. Ihara (eds), The Acid Rain Debate: Scientific, Economic, and Political Dimensions (Boulder: Westview Press, 1985). 676 In fact, Sweden’s major case study for the Conference, led by one of the leaders of international climate science, Bert Bolin, involved acid rain. Cowling, “Acid Precipitation,” 113A, 115A. See Bert Bolin et al, Sweden’s Case Study for the United Nations Conference on the Human Environment: Air Pollution Across National Boundaries (Stockholm: Norstadt and Sons, 1972).

288 Administration considered levying an SO2 tax tied to energy production from coal as early as 1970; internationally, regional agreements between European nations and bilateral negotiations between the U.S. and Canada yielded first bilateral treaties, and eventually the more broadly international United Nations Convention on Long-Range

Transboundary Air Pollution, proposed in 1979 and ratified in 1983.677 As Gus Speth remembers in Red Sky at Morning, the long-distance impacts of these pollutants on agriculture, species and species’ habitats, and water quality challenged environmentalists to think about air pollution as more than just a local, urban problem.678

Legally, however, acid rain represented little more than a variation on a theme, and

American environmental groups sought to tackle the problem primarily by expanding existing American legal mechanisms at home. Because of their impacts on public health

and natural beauty (they reduced visibility), SO2 and NOx (smog) constituted forms of air pollution in their own right within the “antidegradation” framework of the Clean Air Acts

677 Nixon’s experience with SO2 and the SO2 tax serves as a microcosm of his relationship with environmental issues and the environmental movement more broadly. Nixon proposed the tax in his 1970 Presidential Environmental Message, but a series of memos between White House staffer John C. Whitaker and representatives from both the Treasury Department and the Department of the Interior early in 1971 suggest that Nixon had little in the way of a feasible plan for implementing the tax when he made the proposal. A year later, he was hearing it from the environmental community to put his proposal into action; at the same time, he was taking it on the chin from American mining organizations and the rest of the business community for what they saw as excessive regulation. Ultimately, the SO2 tax faltered in large part because the Office of Management and Budget realized that it could be too successful too quickly, and

SO2 might be taxed out of existence before the revenues from the tax could pay for the pro-business programs Nixon would have to use to sell it politically. Memorandum for Ken Cole from John C. Whitaker, January 6, 1971; “Sulfur Dioxide Emissions Charge,” memo for John B. Connally [sic] Jr., Secretary of Treasury, November 11, 1971; John C. Whitaker, Memorandum, November 17, 1970: all in Nixon Presidential Material Project, White House Central Files, Staff Member Office Files, John Whitaker Papers. 678 Sam Hays’s analysis concurs with Speth’s analysis almost to a T. “Acid rain,” Hays wrote in 1987, in a sentence that could have foreshadowed a discussion of global warming (but didn’t), “was a widely shared problem that galvanized cleaner-air advocates across many interests, providing a potential base for extending the concern for air quality far beyond urban centers.” Hays, Beauty, Health, and Permanence, 123. James Gustave Speth, Red Sky at Morning: American and the Crisis of the Global Environment (New Haven: Yale University Press, 2004), 52.

289 of 1967, 1970 and 1977.679 The identifiable impacts of these gases on water quality and plant growth via acid rain allowed environmentalists to capitalize on the provisions of the

Clean Water Act of 1972 as well, increasing their options as they lobbied for lower emissions thresholds and prepared lawsuits against businesses, industry, and government agencies.680 Meanwhile, these same identifiable impacts, once publicized, helped to mobilize grassroots support for litigation and lobbying efforts from environmental constituents concerned about everything from local public health issues to the sanctity of

America’s wilderness areas. Acid rain thus gave environmentalists both legal and political traction on an important large-scale atmospheric problem—a problem the Global

2000 report framed alongside CO2-induced climate change and ozone depletion—without necessitating a wholly new strategy of environmental activism.

Acid rain helped link environmentalists’ ideas about regional air pollution with concerns over large-scale atmospheric change, but it was the degradation of stratospheric ozone that brought environmentalists around to the protection of the global atmosphere as

a whole. Ozone (O3) is an allotrope (that is, a natural structural modification) of oxygen

(O2), present throughout the various layers of the atmosphere. In the lower atmosphere, ozone is both a pollutant and an effective greenhouse gas. High levels of ozone cause respiratory problems in animals and can “burn” the leaves of sensitive plants. In the stratosphere, though, where most of the world’s ozone is created and destroyed, the gas works to absorb ultraviolet energy from the sun, protecting the biosphere below from

679 Hays, Beauty, Health, and Permanence, 160-162. 680 In 1967, lobbyists for coal and coal-using companies managed to secure a raise in the allowable level of

SO2 from .02ppm to .03ppm, but environmentalists soon used the issue of acid rain to undercut this change. In 1981, they argued that the threshold for harm from acid deposition stood at 20 kilograms of sulfate per hectare per year—a rate of deposition that would require standards far below .03ppm. Hays, Beauty, Health, and Permanence, 162-163, 352.

290 potentially damaging wavelengths of light.681

In the late 1960s and early 1970s, concerns over ozone depletion revolved around the potential for water vapor and oxides of nitrogen from high-elevation supersonic transport flights to accelerate natural stratospheric ozone destruction, but concern about ozone depletion did not die with the end of the American SST program. 682 As Paul

Crutzen, Harold Johnston, and others studied the effects of stratospheric NOx on ozone in the mid-1970s, they soon began to realize that other substances that, like NOx, seemed stable in the troposphere, might undergo chemical changes as they interacted with radiation and ozone in the stratosphere.683 In 1974, Sherwood Rowland and Mario

Molina published a groundbreaking paper that investigated the role of halogenated hydrocarbons (also known as halocarbons, molecules of carbon and hydrogen covalently bonded to atoms of the halogen family, including fluorine, chlorine, bromine, and iodine) in the destruction of stratospheric ozone.684 The paper, which earned the two chemists the

1995 Nobel Prize in chemistry, addressed two specific chlorofluoromethanes (CF2CL2

and CFCL2) commonly used as refrigerants and as propellants in aerosol spray cans.

Molina and Rowland also suggested that in addition to CF2CL2 and CFCL2, chlorofluorocarbons, or CFCs, along with other halocarbons like those in the widely used

681 See Denis L. Hartman, Global Physical Climatology (San Diego, CA: Academic Press, 2004), 49, 70- 71, 324-327. 682 For more on SSTs, ozone, and the relationship between atmospheric scientists and the environmental movement in the 1960s and early 1970s, see Chapter 2. 683 P. J. Crutzen, “The Influence of Nitrogen Oxides on the Atmospheric Ozone Content,” Quarterly Journal of the Royal Meteorological Society, vol. 96, no. 408 (December, 1970): 320–325. Summarized in Harold S. Johnston, “Atmospheric Ozone,” Annual Review of , vol. 43 (1992): 4. 684 Halocarbons impact stratospheric ozone through a process called “photodissociation,” wherein certain wavelengths of radiation cause halogens to separate from their associated hydrocarbons. The result is loose halogen molecules—most notably chlorine—that essentially steal the third oxygen atom from ozone, and then relinquish it to another oxygen atom to create an O2 molecule, leaving the halogen to catalyze the process all over again. The stability of halogens, meanwhile, allows them to stay in the stratosphere almost indefinitely, causing indefinite destruction to stratospheric ozone. Mario J. Molina and F.S. Rowland, “Stratospheric Sink for Chlorofluoromethanes: Chlorine Atom-Catalyised Destruction of Ozone,” Nature 249 (June 28, 1974): 810-12. See also See Hartman, Global Physical Climatology, 49, 324-327.

291 DuPont refrigerant Freon, might also catalyze the destruction of stratospheric ozone.

Scientists working in conjunction with UNEP and the WMO soon created projections of future CFC and halocarbon usage, from which they then attempted to extrapolate future ozone depletion.685 By the time the Global 2000 Report to the President laid these issues out for policymakers within the context of the world’s environmental systems in 1980, it was clear that these substances presented a real danger to stratospheric ozone, and, by association, to both the human and non-human denizens of Earth.

As an environmental issue, ozone depletion bridged the gap between regional air pollution problems subject to control within existing legal and political

frameworks—problems like acid rain—and the global issue of CO2-induced climate change. Ozone depletion and global warming shared two important attributes that distinguished them politically from acid rain. First, though Global 2000 expressed

concern over the expanding geographical reach of SO2 and NOx, the two gases’ environmental and public health impacts directly affected the geographically discrete places where either the gas settled and people breathed it in, or where the chemicals fell

as precipitation. CO2 and stratospheric ozone, by contrast, only affected local human and environmental systems through their global action. The ozone “hole”—discovered as early as 1976 but only confirmed in the middle of the 1980s—had a vague geographical range in the Southern Hemisphere, but for the most part, ozone depletion led to increased levels of UV-B reaching the Earth’s surface throughout the globe; it was thus a truly global problem.686 Similarly, though climate change might impact different regions of the

685 Speth, Red Sky At Morning, 92-93. 686 The story of NASA’s “discovery” is an interesting one, told well by Erik Conway in Atmospheric Science at NASA: A History (Baltimore: Johns Hopkins Press, 2008). Consensus among atmospheric chemists held that there had been about a 6% depletion of ozone globally, with few regions significantly

292 globe to varying degrees depending on factors like latitude and existing marginal

precipitation, CO2 itself represented a danger only through its ability to impact the

Earth’s total energy budget.

Second, as Global 2000 underscored, both ozone depletion and global warming existed as problems primarily in the future tense. Most new environmental issues began with the recognition of obvious impacts, usually driven by direct observation. As early as

1852, for example, Angus Smith noticed that the sulfuric acid in the air in Manchester caused the colors of textiles to fade and metals to corrode.687 True concern over acid rain began with studies of the measured acidity of Scandinavian rivers and streams; further investigation led to more robust scientific theory.688 At least until the discovery of the ozone hole, concerns over both ozone and global warming turned this process on its head; scientific theory presaged future impacts, with few measurable data-points to drive the problem home in the present.

There was also a direct political relationship between efforts to protect the ozone layer and the burgeoning fight against anthropogenic global warming. The debate over ozone helped to make the idea of a global convention on climate change politically feasible in the first place. As Alan Hecht and Dennis Tirpak of the EPA wrote in their

depleted beyond that 6% global average. NASA’s satellite measurements generally confirmed this number, until a paper by Joseph Farman appeared that challenged the numbers coming from the Dobson instruments (the instruments that measure ozone) on NASA’s satellites. When scientists at the Goddard Space Flight Center revisited their readings, they realized that the computer program written to handle the data had used a quality-control code designed to flag and essentially throw out data outside of a certain range of Dobson units (below 180). Richard Stolarski of Goddard reexamined the data and realized that the computers had in fact masked a huge “hole” in ozone over Antarctica comprised of a continent-sized region in which ozone levels had dropped below 150 Dobson units. See Conway, 171-173; Sharon L. Roan, Ozone Crisis: The 15 Year Evolution of a Sudden Global Emergency (New York: Wiley and Sons, 1989), 131; Edward A. Parson, Protecting the Ozone Layer: Science and Strategy (New York: Oxford University Press, 2003); 84- 85. See also J.C. Farman, B.G. Gardiner, and J.D. Shanklin, “Large Losses of Total Ozone in Antarctica

Reveal Seasonal ClOx/NOx Interaction,” Nature 315 (May 16, 1985): 207-10. 687 Cowling, “Acid Precipitation,” 111A. 688 Ibid., 113A.

293 summary of the UNFCCC in 1995, “one of the main consequences of the CFC debate was to convey to the general public the concept of a real global environmental threat and the need for domestic and international action to address it.”689 The ozone issue helped to establish better coordination between scientific organizations like the ICSU and National

Academy of Sciences, environmental NGOs like NRDC, the Sierra Club, and

Greenpeace, and international agencies like WMO and UNEP. In a more limited way, it also fostered the type of cooperation between the Department of Energy and the EPA that would become essential in negotiating a framework convention on climate change in the future.690

As Gus Speth argues, the development of the 1987 Montreal Protocol on Ozone

Depleting Substances came to serve as a sort of standard model for international policymaking on global environmental issues more generally in the late 1980s.691 In an ideal world, Speth explains, the process evolves in four stages. First, scientists—in this case Molina, Rowland, Crutzen, and others—identify a problem and conduct research in order to produce an agenda for non-governmental advocacy based on good scientific evidence. Second, after more research and “fact-finding”—run in the case of ozone primarily by UNEP and the WMO—the international community negotiates a broad set of policy goals and agrees upon the political procedures of a framework convention designed to accomplish these goals. The framework convention typically yields a treaty—the 1985 Vienna Convention—enforced through negotiated protocols—most notably the Montreal Protocol—that describe the more specific actions that individual

689 Alan D. Hecht and Dennis Tirpak, “Framework Agreement on Climate Change: A Scientific and Policy History,” Climatic Change, vol. 29, no. 4 (April, 1995): 377. 690 Ibid. 691 Speth, Red Sky at Morning, 91-92.

294 nations must take to continue to accommodate the treaty as the problem changes over time. Third, a series of national-level political campaigns secure the convention’s ratification within individual national governments, who then collectively codify the treaty under the rules of the United Nations. And finally, once in force, the protocols are implemented, monitored, and strengthened depending on their success over time.692 As

Speth writes, “the Montreal protocol is the crowning achievement of global environmental governance.”693 And indeed, it has succeeded remarkably in stemming the tide of stratospheric ozone depletion in the past two decades.694

The Montreal Protocol has served less admirably, however, as a model for

international action on climate change. Unlike the buildup of CO2, ozone depletion had a direct, tangible impact on human health—increases in the risk of skin cancer—that struck a chord with an environmental community already deeply engaged with the fight against cancer and keen on reducing environmental carcinogens through the National Toxicology

Program and other related public health campaigns.695 The National Academy of

Sciences estimated that the continued use of CFCs at 1974 rates could lead to a 14% reduction in stratospheric ozone, and they made the impacts of this kind of reduction abundantly clear. In a world of reduced ozone, “all unshielded cells are highly vulnerable to sunlight and may be killed by relatively short exposure to full sunlight,” the Academy wrote.696 As Global 2000 pointed out, a mere 10% reduction in stratospheric ozone

692 Ibid., 92-94. 693 Ibid., 94. 694 Ibid., 92-94. 695 For the National Toxicology Program, see Hays, Beauty, Health, and Permanence, 354-55. 696 Cited in Global 2000, 265. The origin of the quotation about skin cancer is elusive, though it iis most likely the 1975 CIAP report, National Research Council, “Biological and Medical Effects of Nitrogen Oxide Emissions,” Chapter 3 in Environmental Impact of Stratospheric Flight: Biological and Climatic Effects of Aircraft Emissions in the Stratosphere, Report by the Climatic Impact Committee for the National Academy of Sciences and the National Academy of Engineering (Washington, D.C.: National

295 “appears likely to lead to a 20-30% increase” in skin cancer.697 “Ozone induced changes in UV radiation,” the report noted,

“would change one of the conditions that has almost certainly influenced the

evolution of life on earth so far, and a significant UV increase [from ozone

depletion] can be expected to precipitate a disturbance in the existing balance of life

virtually everywhere on the planet.”698

Whereas climate change remained somewhere between a legitimate environmental concern and a scientific curiosity, environmentalists could treat ozone depletion as an environmental emergency.

Also unlike threats to the stability of the Earth’s climate, artificial threats to stratospheric ozone came from discrete and easily identifiable sources, mostly in the form

of non-essential or replaceable consumer products. Atmospheric CO2, by contrast, came from fossil fuels burned around the world for energy. These fuels stood at the heart of the 20th century world economy, and unlike CFCs, they had few ready replacements—none at a global scale. Aerosol cans and halocarbon refrigerants were easy to vilify, especially for an environmental movement that had cut its teeth on more than a decade of consumer advocacy campaigns. The few large U.S. chemical companies that produced these products gave groups like Environmental Defense and NRDC a target for the type of domestic legal strategies that had served them so well on other environmental issues. As early as 1976, a “Ban the Can” public relations campaign,

Academy of Sciences). Global 2000 instead cites National Research Council, Panel on Nitrates, Nitrates: An Environmental Assessment (Washington, D.C.: National Academy of Sciences, 1978), and the Global 2000 Report also takes information on the links between ozone depletion and skin cancer from National Research Council, Committee on Impacts of Stratospheric Change, Halocarbons : Effects on Stratospheric Ozone (Washington DC: National Academy of Sciences, 1976). 697 Global 2000, 265. 698 Ibid.

296 alongside a legal push from NRDC, led the EPA to pursue a phase-out in the production of aerosol products containing fluorocarbons, effective in December of 1978.699 By the early 1980s, moreover, many of the companies that manufactured CFCs in the first world had already found ozone-friendly substitutes, which they could relatively easily phase in over time to meet the voluntary standards of the Montreal Protocol.

Finally, the existence of the Montreal Protocol itself changed the international political landscape in which global warming advocates hoped to develop a convention on climate change. In creating a consensus on ozone, UNEP executive secretary Mustafa

Tolba and his colleagues had the triple benefit of a certain level of pre-existing national- level political buy-in from around the industrialized world, the eager support of a capable group of concerned scientists, and perhaps most importantly, the element of surprise.

Much of process for creating an international scientific consensus on ozone depletion and translating that consensus into an international legal regime unfolded in an ad hoc manner. International agencies like UNEP and WMO, alongside scientific and environmental NGOs like the ICSU and NRDC, all carved out niches within this new international legal framework as they developed it. The few companies and governments still opposed to regulation on CFCs and other ozone depleting gases in the 1980s found themselves decidedly on the defensive, reacting to these new forms of environmental advocacy as they arose.

During negotiations over the UNFCCC, by contrast, opponents of action on climate change took the initiative, and they were able to do so in part because of their

699 Andre T. McCloskey, “The Stewardship of Planet Earth: Protecting Our Planet Twenty Years of Defending the Environment,” in BIOS, vol. 60, no. 1/2 (March - May, 1989): 40; Environmental Protection Agency, “Government Ban on Fluorocarbon Gases in Aerosol Products Begins October 15 [1978],” EPA Press Release, October 15, 1978, U.S. Environmental Protection Agency, “History,” http://www.epa.gov/history/topics/ozone/01.htm.

297 experiences with the ozone debate. The international environmental community’s success in regulating CFCs put the much more powerful corporate and government bodies opposed to regulations on energy and fossil fuel emissions on notice. The Vienna-

Montreal process—UNEP’s explicit model for the UNFCCC—provided politically powerful energy industry lobbyists and energy-hungry governments with a rough outline of what they could expect to face when it came to global warming. Unlike with ozone, opponents of regulation on greenhouse gases consequently had the opportunity to carve out their own niches alongside environmental NGOs in the negotiation process in order to tip the strategic playing field on climate change in their favor…and they did so. Having watched Tolba and his NGO supporters take ownership of the ozone issue by capitalizing on the momentum generated through the consensus-making process, leaders from energy- dependent and energy-producing industries joined the Reagan and Bush Administrations in undercutting the UNFCCC process both scientifically and politically at every step of the way. Thus while the Vienna Convention and the Montreal Protocol helped put atmospheric change on the international political map, these agreements’ very existence upped the political ante on the climate problem moving forward. “Politics caught up with ozone,” noted one of Tolba’s key advisors during the UNFCCC process, but “climate was born in politics.”700

Mechanisms of Knowledge: Villach 1985 and the IPCCC

Nothing reflected the political limitations of the Vienna-Montreal process as a model for climate change more clearly than the relationship between science and policy

700 Interview with Peter Usher by Shardul Agrawala, Bonn, Germany, March 4, 1997, as cited in Shardul Agrawala, “Context and Early Origins of the Intergovernmental Panel on Climate Change,” Climatic Change, vol. 39, no. 4 (August, 1998): 614.

298 during the development of the UNFCCC. As with ozone, the process of developing an international legal regime for climate change revolved around the production and validation of scientific knowledge. Initially, scientists sought to develop an international consensus on climate change through familiar processes of scientific assessment similar to those that had proved effective on ozone and acid rain. As scientists, environmental

NGOs, and American and United Nations environmental agencies soon recognized, however, the form of independent scientific consensus that fostered political action on

ozone proved insufficient as a jumping-off point for international regulations on CO2 and other GHGs. Governments and corporations understood that international scientific consensus on climate change presaged international action on the issue, and they insisted on maintaining some level of control not just over the political process, but also over the science. The U.S. government, representatives of the nation most responsible for increases in GHGs and thus perhaps the most important party in climate change negotiations, refused to accept a consensus report unless government scientists—and, by association, the Reagan Administration—had an active role in producing it. Without backing from powerful governments like the U.S., NGOs and U.N. agencies found that scientific consensus yielded very little real political action. If Mustafa Tolba and his colleagues hoped to replicate their success with ozone by using science to support an international convention on climate change, they first needed to establish political buy-in on the science itself. The mechanism they developed for this purpose was the

Intergovernmental Panel on Climate Change, the IPCC.

In the early 1980s, climate scientists already had a number of mechanisms for creating independent scientific consensus, both domestically and internationally.

299 Domestically, the most prestigious and important of these was the National Research

Council of the National Academy of Sciences. Until at least the late 1970s, the NAS had served as the definitive arbitrator of scientific disagreements surrounding climate change.

Largely at the behest of the federal government, by 1985 the NAS had produced a handful of major assessments of the various relevant components of climate, including the 1977 Energy and Climate, the 1979 Charney Report, and the 1983 Nierenberg assessment.701 In the 1980s, however, scientists and environmentalists, working with

Congressional Democrats, began to use climate science and the issue of climate change to undermine the Reagan Administration’s energy and defense policies.702 Along the way, both politically active scientists and political actors themselves from both sides of the climate issue also began to exercise more and more control over the NAS consensus- making process. The dual EPA and NAS reports of 1983 underscored this politicization, and as a result, National Academy of Sciences Reports carried only so much weight as accurate assessments of the state of the science in the 1980s.

Abroad, meanwhile, the International Council of Scientific Unions continued to work with the WMO and UNEP to promulgate the global environmental monitoring and research efforts established over the previous three decades. Following the World

Climate Conference of 1979, the ICSU and WMO hosted a series of conferences on climate change in Villach, Austria, designed to provide a thorough international

701 See National Academy of Sciences, Geophysics Research Board, Energy and Climate: Studies in Geophysics, (Washington, DC: National Academy of Sciences, 1977); National Academy of Sciences, Climate Research Board, Carbon Dioxide and Climate: A Scientific Assessment (Jule Charney, Chair) (Washington, DC: National Academy of Sciences, 1979); National Research Council, CO2/Climate Review Panel, Carbon Dioxide and Climate: A Second Assessment, (Washington, DC, National Academy of Sciences, 1982); National Academy of Sciences, Carbon Dioxide Assessment Committee, Changing Climate, (Washington, DC: National Academy of Sciences, 1983), also known as the “Nierenberg Report” for it’s chairman, William Nierenberg. See also National Defense University, Climate Change to the Year 2000: A Survey of Expert Opinion, (Washington, DC, National Defense University, 1978). 702 See Chapter 5.

300 assessment of the problem. Operating under the auspices of the ICSU, however, these assessments self-consciously avoided the political implications of climate change. They focused instead on updating the environmental monitoring systems the ICSU created in the 1960s and 1970s and establishing better methods of data analysis and climate modeling. Their recommendations revolved around international research efforts, and generally shied away from international politics.

Until 1985. Unlike its predecessors, the consensus report of the last of these

Villach conferences, held in October of that year, contained a somewhat vague but nevertheless striking set of recommendations for moving forward with policy on climate

change. If CO2 and other GHGs were to continue to accumulate in the atmosphere, the report ominously warned,

“in the first half of the next century a rise of global mean temperature would occur

which is greater than any in man’s history…Many important economic and social

decisions are being made on long term projects…based on the assumption that past

climatic data…are a reliable guide to the future. This is no longer a good

assumption since the increasing concentrations of GHGs are expected to cause

significant warming of the global climate in the next century.”703

In response, the Villach authors contended, “scientists and policymakers should begin active collaboration to explore the effectiveness of alternative policies and

703 “Statement” in World Meteorological Organization, Report of the International Conference on the Assessment of the Role of Carbon Dioxide and of Other Greenhouse Gases in Climate Variations and Associated Impacts, Villach, Austria, 9-15 October, 1985, WMO No. 661 (Geneva: World Meteorological Organization, 1986). Also available online from the ICSU as SCOPE 29 at http://www.icsu- scope.org/downloadpubs/scope29. Quoted in Jaeger, Developing Policies, 1-2. Shardul Agrawala gives a readable and detailed, though not particularly analytically critical account of the creation of the IPCC in “Context and Early Origins.” He quotes WMO 1985 on pg. 608. See also Hecht and Tirpak, “Framework Agreement”; Michael Oppenheimer, “Developing Policies for Responding to Climate Change,” Climatic Change, vol. 15, no. 1 (1989): 1-4; Stephen Schneider, Science as a Contact Sport, 124-126.

301 adjustments.”704 Spearheaded by Swedish meteorologist and Stockholm University professor Bert Bolin (a long-time advocate for better climate research), a group of scientific leaders at Villach recommended that the ICSU, WMO, and UNEP modify the assessment process itself in order to begin providing information and recommendations for an international treaty that would mitigate what Bolin and many of his ICSU colleagues characterized as an increasingly dire global environmental problem.

In 1986, Bolin formed a small, informal group of international climate scientists called the Advisory Group on Greenhouse Gases (AGGG) to do just that. Comprised of

Bolin and two members from each of three organizations—UNEP, WMO, and

ICSU—AGGG took as its mandate to establish a task force to look into the potential impacts and policy responses to climate change, and to “initiate if necessary, consideration of a global convention.”705 In 1987, with a grant from the Rockefeller

Brothers Fund, as well as support from the Environmental Defense Fund via Michael

Oppenheimer and from George Woodwell and his Woods Hole Research Center (among other institutions), AGGG launched a two-part conference specifically focused on developing national and international policies for responding to climate change. Held in

Villach, Austria under the auspices of Stockholm’s Beijar Institute in early October, the first conference laid out the global and regional impacts of greenhouse warming and addressed “technical, financial, and institutional options for limiting or adapting to climatic changes.”706 The second conference, held a month later in Bellagio, Italy, used the initial assessment as a platform for proposing specific policies for mitigating and adapting to climate change in various regions, and suggested a set of international

704 “Statement,” WMO, International Assessment, 1985. 705 Ibid. Also cited in Agrawala, “Context and Early Origins,” 609. 706 Jaeger, Developing Policies, i.

302 institutional arrangements that could help in the implementation of these policies.707

For climate scientists, Villach and Bellagio marked a new, more aggressive approach to climate policy. The group recommended setting specific “long term environmental targets, such as the rate of temperature or sea level change…based on observed historic rates of change that did not put stress on the environment” as a platform for international decision-making.708 They suggested that international policymakers set goals that would limit sea level rise to between 20 and 50mm per decade and global mean temperature rise to 0.1ºC per decade. The group stated flatly that limiting global warming “could only be accomplished with significant reductions in fossil fuel use,” and they weighed the costs of these reductions against the potential costs of doing nothing.709

Most forcefully, they emphasized the immediate need for international regulatory mechanisms for limiting GHG emissions, and particularly for “an agreement on a law of the atmosphere as a global commons” or “a convention along the lines for that developed for ozone.”710

Bolin and the AGGG brought their activist momentum into the 1988 World

Conference on the Changing Atmosphere in Toronto, Canada, a mixed scientific and political meeting sponsored by former Canadian Meteorological Service head Howard

Ferguson that dealt with both climate change and ozone depletion.711 As they had in

Villach and Bellagio, the scientists involved in the Toronto conference again emphasized the need for specific targets derived from past records of environmental change, and

707 Ibid. 708 Ibid., v. 709 Ibid., v, 22, 27. 710 Ibid., v. The need for international regulatory mechanisms recurs through the report. 711 Agrawala, “Context and Early Origins,” 610. See World Meteorological Organization, Proceedings of the World Conference on the Changing Atmosphere: Implications for Global Security, Toronto, Canada, June 27–30 (Geneva: WMO/UNEP, 1988). WMO Doc. 710 (1989).

303 again they called for an international legal framework in which to secure these environmental goals. A group of energy experts at the Toronto conference called for a

20% reduction of global carbon-dioxide emissions from 1988 levels by 2005—a stance that the conference ultimately adopted.712 The scientists involved in the Toronto

Conference spoke primarily as individuals or members of NGOs, separate from their nations or specific academic institutions, and the event was as much about publicity as it was about science. It drew a large and moderately powerful crowd, including some heads of state, but it had no official relationship to the forthcoming UNFCCC, and ultimately its recommendations carried little political clout. Still, the Toronto meeting reinforced a growing sentiment within the international scientific community that scientists had to help translate the alarming scientific evidence that the world was warming into real, tangible international political action.

For Bolin and the panoply of scientific and environmental NGOs present in

Toronto, Villach 1985 offered an authoritative consensus on the science of climate that could support the bold targets presented at Villach, Bellagio, and Toronto in 1987 and

1988.713 Tolba and his United Nations colleagues agreed, and with good reason. The

Villach 1985 conference’s participants came from universities and scientific organizations in 29 countries in both the developed and developing worlds, and their conclusions, which generally matched those of the most prestigious scientists working in the United States and England, reflected the state of the art in climate science. The

Villach report expressed plenty of uncertainties, but then so too had reports on ozone depletion in the fact-finding endeavors leading up to the 1985 Vienna Convention

712 Agrawala, “Contexts and Early Origins,” 610. 713 Jeager, Developing Policies, 2.

304 contained a healthy dose of uncertainty. By 1987, Tolba had already begun planning an international convention for climate change—modeled after the ozone convention—and he used Villach 1985 as his scientific support.714 NGOs like the AGGG and U.N. agencies like UNEP had little chance of implementing aggressive policies on energy and emissions without government backing, but scientists and some U.N. leaders felt that

Villach sent a scientific message sufficiently clear and definitive to garner support even from relatively conservative administrations in the U.S. and the U.K. Indeed, immediately following Villach 1985, Tolba appealed to Secretary of State George Shultz, a representative of the largest producer of GHGs and the most important player in any potential climate convention, urging the U.S. to take “appropriate” policy actions in preparation for a new convention.715

Few national governments were ready to accept Villach 1985 as the definitive word on climate, however, least of all the United States. Domestic government agencies continued to disagree over the nature and magnitude of the problem, and the

Administration had no reason to support bold action. Perhaps more importantly, neither the DoE or the EPA—the two main agencies responsible for policy on climate change—had a place at the table in Villach, and the State Department felt the U.S. could not accept an international scientific consensus that did not involve U.S. government scientists who could speak for the interests of the Administration. As an alternative, the

National Climate Program’s policy board proposed the development of a new, intergovernmental body to oversee a comprehensive government-led assessment of climate science that could serve as the basis for a framework convention on climate

714 Agrawala, “Context and Early Origins,” 612. 715 Agrawala, “Context and Early Origins,” 612; Hecht and Tirpak, “Framework Agreement,” 380.

305 change within the United Nations.716 UNEP and WMO soon parlayed this proposal for a new intergovernmental consensus-making mechanism into the Intergovernmental Panel on Climate Change. They charged the new organization with the express mission of creating a consensus on science and policy that could serve as the foundation for the

United Nations Framework Convention on Climate Change.

Officially commissioned in 1988 and chaired by Bert Bolin, the IPCC rehashed much of the same material discussed at Villach, but under a slightly different and more transparent—but also overtly political—assessment structure. Peopled primarily by scientists from government agencies and mid-level diplomats, the IPCC consisted of three main working groups: Working Group 1 dealt exclusively with the state of the physical science of climate change; Working Group 2 discussed the potential economic, social, and environmental impacts of the phenomenon; and Working Group 3 provided a set of possible strategies for responding to the threat of climate change.717 UNEP and the

WMO asked for reports from each group by 1990, at which point they planned to meet as

716 The Reagan Administration stood to benefit from a new, intergovernmental scientific assessment of climate change in a number of ways that involved both domestic and international politics. Throughout the 1980s, Democrats in Congress and scientists at the National Climate Program—an executive level agency operating as “an outpost in enemy territory,” as one scientists put it—continued to put pressure on the President to formulate some sort of national strategy for dealing with climate change. In 1986, then Senator Joe Biden introduced an initiative mandating that the President formulate such a strategy and deliver it to Congress, along with legislative recommendations, within a year. The initiative soon became the Global Climate Protection Act of 1987. A new, international scientific assessment mechanism could at once satisfy the immediate demands of the Biden initiative by demonstrating the Administration’s good faith commitment to finding scientific consensus on the issue and buy time for executive agencies—the EPA, Department of Energy, and State Department—to get their ducks in a row should a treaty on climate become necessary. The development of a new scientific assessment process also necessarily involved official international diplomatic negotiations, which promised these executive agencies—and the State Department in particular—renewed control over what had in the 1980s become largely a Congressional issue. John Rich to Senator Biden, “The Biden Initiative on Global Warming,” September 23, 1986, Working Files of Peter W. Galbraith, Committee on Foreign Relations, 96-102nd Congress, RG 46 Record of the U.S. Senate, Box 34, National Archives and Records Administration (NARA 1), Washington, D.C.; Agrawala, “Context and Early Origins,” 613-614; Hecht and Tirpak, “Framework Agreement,” 381. 717 In addition, a smaller, less formal fourth working group tackled the specific concerns and problems of developing nations, and an administrative bureau oversaw the larger process as well. See Hecht and Tirpak, “Framework Agreement,” 385.

306 a larger body to compile the reports into a single, authoritative assessment and to hash out the details of an executive summary of that assessment.

The key to the IPCC was its intergovernmental character. Unlike at Villach,

Bolin and his UNEP/WMO colleagues self-consciously sought to engage national representatives in the consensus-making machinery of the IPCC. The IPCC consensus was to be more than an agreement between scientists or NGOs; it was to be an agreement between governments. In part, the intergovernmental nature of the new assessment came in response to the U.S. Department of Energy’s criticisms of the nongovernmental nature of Villach; indeed, the Administration insisted that the consensus process involve official government scientists.718 But independent of the DoE, Bolin and Mustafa Tolba also recognized the importance of establishing a process that gave international political actors ownership over the issue of climate change. By participating in the IPCC, even skeptical governments—the U.S., the Soviet Union, and many countries in the developing world—tacitly agreed that the problem merited some sort of international solution. Moreover, once the IPCC released its conclusions, participating nations theoretically had little recourse to objections about the basic facts of the matter when it came time to negotiate a framework convention on climate change. Scientists and U.N. leaders thus traded a certain amount of control over the process of producing consensus

718 The DoE’s objections to Villach ran the gambit from its nongovernmental structure to its actual scientific conclusions. Their qualms with Villach stemmed from two main sources, however. First, the conservative Reagan Administration either didn’t believe or didn’t want to admit that the science of climate change could be expressed with enough certainty to justify possibly expensive international policy. Second, the DoE itself ran a major assessment of climate change in the 1980s, and in the end the agency had to compete with the UNEP/WMO/ICSU project for scientific participants, which undermined its own consensus. The DoE also attempted to convince the WMO to incorporate its assessment into Villach 1985, which the WMO showed little interest in. Agrawala, “Context and Early Origins,” 614.

307 on the science of climate change in return for an implicit political commitment to cooperation on some sort of international climate policy mechanism in the future.719

The IPCC and the New Politics of Consensus

The overt ties between the IPCC and the forthcoming UNFCCC not only helped establish a level of political buy-in on future climate change policy; these ties also raised the political and economic stakes of the IPCC’s assessment of climate science itself.

Whereas Villach represented another in a series of warnings from the scientific community, Bolin and his WMO and UNEP colleagues meant the IPCC to serve as a politically negotiated scientific baseline for real, binding international laws and policies.

Both scientists and politicians already recognized that any serious effort to mitigate CO2 and other GHGs involved major changes in national-level energy and land use strategies—two sectors fundamental to economic growth in developing and industrial nations alike. But they also recognized that the strength and character of any international legal framework for dealing with climate change depended largely on the level of certainty and concern expressed in the IPCC assessment. Consequently, debate over the language and details of the IPCC’s first assessment report almost immediately began to serve as the front line of a larger battle over a future international legal regime—the UNFCCC—that many believed could have a major impact on national

719 Stephen Schneider remembers a conversation with Bolin on the matter in Science as a Contact Sport in which Bolin spelled out his argument. Skeptical of another assessment that would simply delay action on global warming, Schneider complained that climate science had become bogged down in assessment work and that a new assessment process would only further divert resources away from original research and add little to the myriad national reports of the 1980s. “But how many of those [assessments] are convincing people in India or Indonesia or developing countries who don’t trust the science that comes out of it?” Bolin asked. “Will it be possible to have climate policy without having a scientific group in which various countries in the world have some political ownership?” Schneider, Contact Sport, 125.

308 economies and on the world economy as a whole.

The IPCC was well in the works by 1988, but a series of political, scientific, and environmental events in 1987 and 1988 heightened the urgency and political significance of the assessment process, especially for the United States. Throughout the mid-1980s,

Al Gore and his fellow Congressional Democrats continued to push the Reagan

Administration to take steps to begin dealing with the problem of climate change. The

Administration, though it had softened its position on funding for climate research, continued to treat global warming by and large as a non-issue. In 1986, then Senator Joe

Biden introduced an initiative mandating that the President commission an executive- level task force to devise a strategy for dealing with climate change—a strategy the

President was meant to deliver to Congress within one year.720 The initiative became the

Global Climate Protection Act of 1987, which the President signed.721 The bill in the end required very little real commitment from the Administration, but it demonstrated an expanding Congressional interest in the issue that raised the domestic political profile of the State Department’s international negotiations with UNEP and the WMO leading up to the IPCC. It also led to another series of Congressional hearings on the issue, which helped to keep global warming in the news and thus on the public agenda.722

720 John Rich to Senator Biden, “The Biden Initiative,” September 23, 1986. 721 Rafe Pomerance, “The Dangers from Climate Warming: A Public Awakening,” in The Challenge of Global Warming, edited by Dean Edwin Abrahamson (Washington, D.C.: Island Press, 1989), 259-69. Pomerance makes specific reference to the Global Climate Protection Act on 264-5. See also Spencer Weart, “Government: The View From Washington, D.C.” in The Discovery of Global Warming: A Hypertext History of how Scientists Came to (Partly)Understand What People are Doing to Cause Climate Change, http://www.aip.org/history/climate/Govt.htm. 722 See, for example, U.S. Congress, Senate Committee on Environment and Public Works, Hearing, Global Warming, 99th Congress, 1st Session, December 10, 1985 (Washington, D.C.: U.S. Government Printing Office, 1986); U.S. Congress, Senate Committee on Environment and Public Works, Hearings, Ozone Depletion, the Greenhouse Effect, and Climate Change, 99th Congress, 2nd Session, June 10 and 11, 1986 (Washington, D.C.: U.S. Government Printing Office, 1986); U.S. Congress, House Committee on Interior and Insular Affairs, Hearings, Implications of Global Warming for Natural Resources, 100th Congress, 2nd

309 Nothing put climate in the public eye more than the environmental events of the following year, however. In the summer of 1987, just after the publication of the World

Council on Environment and Development’s alarming assessment of the global environment, Our Common Future (more commonly called the Brundtland Report), a heat wave hit the eastern U.S., followed late in the year by the onset of a severe drought in the major farm states of the mid-west. May, June, and July of 1988 brought even more heat to the eastern U.S. and Canada in what one author called “one of the most intense combinations of drought and heat since the Dust Bowl of the 1930s.”723 Crops in the

Great Plains began to fail. The Mississippi River hit a record low water level, compromising the river’s navigability. High winds and severe drought turned a number of small fires in Yellowstone National Park into the largest conflagration in the park’s recorded history, affecting 36% percent of the land within its boundaries.724 Abroad, a bad hurricane season racked the Caribbean, floods ruined crops, killed livestock, and displaced millions of people in Bangladesh, and droughts in China and the Soviet Union threatened food supplies around the world.725 When WRI’s Andrew Maguire told a

Senate Subcommittee on Environmental Pollution in 1986 that “mankind’s activities are changing the atmosphere in ways that could profoundly affect the habitability of the

Session, September 27th in Washington, D.C. and October 17th in San Francisco, CA, 1988 (Washington, D.C.: U.S. Government Printing Office, 1989). 723 William K. Stevens, The Change in the Weather: People, Weather, and the Science of Climate (New York: Dell Publishing, 1999), 129. 724 For a good resource on the ecological impacts of the Yellowstone fires, see Linda L. Wallace (ed), After the Fires: The Ecology of Change in Yellowstone National Park (New Haven: Yale University Press, 2004). See also Stephen Pyne, “Burning Questions and False Alarms about Wildfires at Yellowstone,” Forum for Applied Research and Public Policy, vol. 4, no. 2 (Summer 1989): 31-39; Norman L. Christiansen et al., “Interpreting the Yellowstone Fires of 1988,” Bioscience, vol. 39, no. 10 (Fire Impact on Yellowstone, Nov., 1989): 678-685. 725 See H. Brammer, “Floods in Bangladesh: Geographical Background to the 1987 and 1988 Floods,” The Geographical Journal, vol. 156, no. 1 (March, 1990): 12-22; Kirkpatrick Sale, The Green Revoluton: The American Environmental Movement 1962-1992 (New York: Hill & Wang, 1993), 71-72.

310 Earth,” his comments echoed a chorus of similar sentiments from scientists and politicians that had begun at least a decade before.726 In the midst of one of the warmest and wildest weather years on record, the media began to ask whether those changes had arrived.727

With the help of Colorado Senator Tim Wirth, James Hansen of NASA’s Goddard

Institute for Space Studies took advantage of the weird weather to add urgency to what was already a growing public concern. Wirth scheduled a hearing on global warming for

June 23, 1988—in the heart of the summer, typically a slow time for Congress—and invited Hansen and GFDL’s Sukiro Manabe to testify. The temperature in Washington,

D.C. that day reached 101ºF.728 Speaking as a “private citizen on the basis of scientific credentials” rather than as an official NASA employee, Hansen told the Senate Energy and Natural Resources Committee—and the media—that he was 99% sure that the world was warming; that scientists could ascribe that warming to the greenhouse effect, and, implicitly, to fossil fuels, with a “high degree of confidence”; and, perhaps most controversially among scientists, that the impacts of global warming could already be detected in atmospheric models. Hansen made it clear that no individual warm day or heat wave or even series of heat waves provided proof of the larger, longer-term phenomenon of global warming, but alongside Manabe, George Woodwell, and Michael

Oppenheimer, he suggested that the warm, unpredictable weather and frequent severe storms of 1987 and 1988 foretold of the conditions of the climatic regime to come.729 As

Hansen most famously told Phillip Shabecoff of the New York Times, “it’s time to stop

726 U.S. Congress, Ozone Depletion, the Greenhouse Effect, and Climate Change, 36. 727 “The Greenhouse Effect? Real Enough,” New York Times, June 23, 1988. 728 Stevens, The Change in the Weather, 129; Conway, Atmospheric Science at NASA, 234. 729 Philip Shabecoff, “Sharp Cut in Fossil Fuels is Urged to Battle Shift in Climate,” The New York Times, June 24, 1988.

311 waffling so much and say that the evidence is pretty strong that the greenhouse effect is here and is affecting our climate now.”730

The media response helped turn global warming—and the environment more broadly—into a major political issue in the 1988 presidential election. Over the next year, climate change appeared on the front page of nearly every major newspaper, and magazines like TIME, Newsweek, and even Sports Illustrated ran cover stories on global environmental problems.731 TIME, which had dubbed Ronald Reagan “man of the year” in 1980, modified the award in order to name the “Endangered Earth” “Planet of the

Year” in 1989.732 Distancing himself from the Reagan Administration’s reactionary environmental policies, Vice President George Bush declared himself “the environmental candidate” during the Presidential primaries (a bold statement considering that Gore was also in the field at the time).733 He specifically tagged global warming as a new and

730 Shabecoff, “Sharp Cut,” 6/24/88. Hansen’s “stop waffling” statement is quoted without irony in most literature on the history of global warming. And for good reason; Hansen’s statement had a great impact on how the press and the public saw the issue of global warming. The phrase itself, however, is not exactly definitive, juxtaposing the imperative “stop waffling” with “the evidence is pretty strong,” not exactly an unequivocal statement. More problematic, perhaps, is the “white knight” role ascribed Hansen in journalistic accounts of the history of climate science. In The Change in the Weather, for example, William K. Stevens paints Hansen’s statement as the “one dramatic stroke” that propelled climate change to “the front burner of international politics” (pg. 133). Spencer Weart and Erik Conway tell the story more carefully, but both still position Hansen at the center of the narrative (Conway’s book is about NASA, so he should be forgiven). But while Hansen’s statement—and his science—was in fact an important catalyst for public interest in climate change, it bears repeating that politicians like Al Gore and Paul Tsongas, environmentalists like Gus Speth, Rafe Pomerance, and Dan Lashof, and scientists like George Woodwell, Stephen Schneider, Bert Bolin, and the entire cast of the WMO and NAS meetings of the early 1980s had maintained nearly the same position—minus the 99% certainty and the claim that warming was presently detectable, which were quite controversial ideas in the scientific community—for a full decade before Hansen testified for Wirth. Indeed, Hansen himself had made similar comments as early as 1982. It was the context rather than the substance of Hansen’s statement that made politicians and journalists pay so much attention, and Hansen in large part had Tim Wirth to thank for that. See Weart, Discovery, 155-157; Conway, Atmospheric Science at NASA, 233-236; Fleming, Historical Perspectives on Climate Change, 134-5. See also Interview of Jim Hansen by Spencer Weart at NASA Center in New York, November 27, 2000, Niels Bohr Library & Archives, American Institute of Physics, College Park, MD USA, http://www.aip.org/history/ohilist/24309_2.html; Mark Bowen, Censoring Science: Inside the Political Attack on Dr. James Hansen and the Truth of Global Warming, New York: Plume, 2008. 731 Hecht and Tirpak, “Framework Agreement,” 384. 732 “Planet of the Year, Endangered Earth,” cover, TIME, January 2, 1989. 733 Sale, The Green Revolution, 72.

312 important environmental challenge, and vowed to counter the greenhouse effect with “the

White House effect.”734

Despite the increasing visibility of the issue, however, once in office, Bush backed off of his commitment to the fight against climate change. The Administration was slow to appoint a science advisor, and Bush received his primary briefings on the issue from

William Nierenberg, then working for the conservative Marshall Institute, which had recently published a report casting doubt on greenhouse warming.735 Following

Nierenberg’s lead, the Administration—and particularly White House Chief of Staff John

Sununu—began to play up the uncertainties of atmospheric models and the climate system itself.736 The President took what historian Tim Walker has called an “America- first, business-first” approach to formulating policy on global warming. The

Administration stressed the “likelihood and severity of economic risks that would be posed by regulation, while downplaying those that might be posed by ‘potential’ climate change.”737 Thus, despite unprecedented public interest and domestic political backing

734 Hecht and Tirpak, “Framework Agreement,” 383; Conway, Atmospheric Science at NASA. Bush made similar comments on a number of occasions; The New York Times quoted him a campaign speech in Erie Metropark, Michigan, on August 31, 1988. “George Bush: From the Text of a Speech Delivered Aug, 31 in Erie Metropark, Mich,” The New York Times, September 24, 1988. See also “The White House Effect. (global warming),” The Economist (US), February 3, 1990. 735 Naomi Oreskes and Eric Conway provide an in-depth analysis of certain scientists’ efforts to quash the issue in the 1980s and early ‘90s in Merchants of Doubt: How a Handful of Scientists Obscured the Truth on Issues from Tobacco Smoke to Global Warming (New York: Bloomsbury Press, 2010). Their work on climate deniers is well-argued and convincing, and has inspired many of the structural questions of this dissertation. See also Oreskes et al, “From Chicken Little to Doctor Pangloss”; Conway, Atmospheric Science at NASA, 238. 736 See William K. Reilly, “Breakdown on the Road from Rio: Reform, Reaction and Distraction Compete in the Cause of the International Environment. 1993-1994” (Arthur and Frank Payne Lecture, Stanford University, Stanford, CA, March 2, 1994), as cited in Hecht and Tirpak, “Framework Agreement,” 402. 737 Walker suggests that the Bush Administration’s approach to environmental problems in general reflected an “economic precautionary principle,” a sort of business-minded inverse of the environmental precautionary principle wherein any environmental regulation must be shaped first and foremost to avoid risks to the economy—and above all to the American economy. The Administration referred to this as a “no regrets” environmental policy. It is effective in the sense that like third wave environmentalism, it privileges policies that increase both economic efficiency and environmental protection—“no-brainers” that easily gain political support. It is not, however, a way to create dramatic change or tackle difficult

313 for climate change, the Bush Administration, like most other national governments, entered international negotiations on climate change in the late 1980s focused primarily on protecting national economic interests. Just as they did at home, within the international community the Administration used doubts about the science of global warming as a first line of defense against aggressive climate change policy. They found a vehicle for these doubts in the IPCC.

For the U.S. and foreign governments alike, climate science served as a surrogate for climate politics at nearly every level of discussion in the first IPCC, beginning with the leadership and structure of the body itself. UNEP and the WMO allocated leadership of the working groups largely based on the political importance of the nations represented.738 John Houghton of the U.K. headed Working Group 1 on basic science; the USSR’s Yuri Izrael took over Working Group 2 on impacts; and a U.S. State

Department official, Fred Bernthal, served as the chief of the politically sensitive

Working Group 3 on potential responses. Each chairperson worked with co- and vice- chairpersons from various other nations so that each Working Group represented a panoply of different and often competing economic and political interests. Bernthal, for example, worked on policy responses with representatives from China, Canada, Malta, the Netherlands, and Zimbabwe; Houghton worked with representatives from Senegal and Brazil.739 From 1988 to 1990, every step of the assessment process, from decisions about what specific questions to address and what scientists to consult to discussions

decisions about global problems like climate change. Tim E. Walker, “Promoting Air Quality, Resisting Climate Change: The George H.W. Bush Administration and the Economic Precautionary Principle,” (paper presented to the Annual Conference of the American Society of Environmental Historians, Boise, Idaho, March 15, 2008). 738 Agrawala, “Context and Early Origins,” 617. 739 Hecht and Tirpak, “Framework Agreement,” 385.

314 about what verbiage to use in presenting the group’s findings, reflected ongoing negotiations between these working group chairs, co-chairs, and vice chairs.

Governments were not the only interested parties involved in the IPCC process.

Once in session, Working Group meetings also included scientific “observers” representing a variety of NGOs. For environmentalists, the organization and influence of their counterparts working for business and industry was an eye-opener. In a May 1990 meeting of Houghton’s Working Group 1 in Berkshire, England, for example, Jeremy

Leggett, a Royal School of Mines geologist turned Greenpeace activist, remembers sitting in the back row alongside Dan Lashof of NRDC and “eleven scientists from the oil, coal and chemical industries, including two from Exxon, one from Shell and one from BP.”740 These observers influenced the IPCC’s assessment in two main ways. First, because the IPCC defined these observers’ roles only loosely, they participated relatively openly in scientific discussions alongside government scientists, especially when those discussions dealt with the wording of the summaries of the Working Group reports.741 In

Working Group 1, scientists working for environmental organization like Leggett and

Lashof consistently pushed the IPCC to spell out explicitly the worst-case scenarios of climate change for policymakers in order to emphasize the full range of risks posed by global warming. Scientists representing the interests of the oil, coal, and gas industries, many of them working for politically conservative, energy industry-funded thinktanks committed to free market ideals like the Global Climate Coalition, the Marshall Institute, and the Global Climate Council, sought to downplay these worst-case studies and emphasize the inherent uncertainties of atmospheric modeling and the myriad, poorly-

740 Jeremy Leggett, The Carbon War: Global Warming and the End of the Oil Era (New York: Routledge, 2001), 3. 741 Leggett, The Carbon War, 3.

315 understood complexities of the global climate system.742

Just as importantly, nongovernmental scientists from both industry and environmental groups also sought to establish relationships with representatives of foreign governments in order to influence these governments’ positions on controversial aspects of the IPCC reports. Here they acted essentially as lobbyists. Again, the two groups focused their respective efforts on attacking or upholding the validity of models and the certainty of the science more broadly.743 And again, environmental NGOs found themselves outnumbered and overmatched by representatives from the energy industry particularly effective in the role.

Negotiations over the specifics of the IPCC consensus only intensified as the

742 Leggett and his colleagues in the environmental community soon began to refer to this group of independent scientists and conservative thinktanks pejoratively as the “Carbon Club.” The most important of these organizations was the Global Climate Coalition, a sort of “umbrella organization for the oil, coal, and auto industries’ response to the global warming issue.” On its board sat representatives from the American Petroleum Institute, Amoco, Arco, Phillips, Texaco, Dupont, Dow Hydrocarbons, Shell, and BP, as well as the Association of International Automobile Manufacturers and the Motor Vehicle Manufacturers Association. The American Electric Power Service Corporation, the American Mining Congress, the Edison Electric Institute, and the National Coal Association also had places on the GCC board. Leggett, The Carbon War, 10-11. 743 The focus on models as “boundary objects” in the climate debate has attracted a wealth of interesting science studies scholarship in the last two decades. As Stephen Bocking writes, “models function as kind of a common language for the scientific disciplines involved in climate research, and also between scientists and society—in effect, a common ground constructed out of all the disparate forms of knowledge about climate.” In their 1998 article, Jeroen van der Sluijs, Josee van Eijndhoven, Simon Shackley, and Brian Wynne demonstrate how scientists themselves have actually decided to discard outliers in the model results used for the executive summaries of consensus reports like Villach 1985 and the IPCC Working Group 1 report in order to protect the communicative power of models in providing quantitative constants—“anchoring devices,” as they call them—for policymakers. Scientists building consensus have maintained the 1.5-4ºC range for the doubling of CO2 identified by the 1979 Charney Report despite model scenarios in which the temperature range for a doubling of CO2 is higher than the range presented in the Charney Report. In “science for policy” situations, the Charney range represents an “anchor” not just of scientific certainty, but of scientific stability. The full reports of assessments like Villach and Working Group 1 have included top-end outliers, but in executive summaries meant for policy-makers, the Charney range persists. Stephen Bocking, Nature’s Experts, 114; Jeroen van der Sluijs, Josee van Eijndhoven, Simon Shackley, and Brian Wynne, “Anchoring Devices in Science for Policy: The Case of Consensus around Climate Sensitivity,” Social Studies of Science, vol. 28, no. 2 (April, 1998): 291-323. See also Simon Shackley and Brian Wynne, “Representing Uncertainty in Global Climate Change Science and Policy,” Science, Technology, and Human Values, vol. 21, no. 3 (Summer, 1996): 275-302; David H. Guston, “Boundary Organizations in Environmental Policy and Science: An Introduction,” Science, Technology, and Human Values, vol. 26, no. 4 (Autumn, 2001): 399-408.

316 working groups presented their reports to the IPCC plenary. In the working groups, scientists representing nations and political interests fought to define the boundaries of the information the IPCC would provide to policymakers. In the plenary, government representatives—including some heads of state—focused on protecting their nations’ interest by adjusting the specific language in which the international community would present that body of knowledge to the world. European states fought to maintain the original language of the Working Group reports in the final summaries, but the U.S., the

Soviet Union, and a number of other countries took the opportunity to water down statements on the impacts of climate change agreed on by Working Group 2 and to highlight the uncertainties involved in the basic science covered in Houghton’s Working

Group 1 summary (Working Group 3, led by the U.S., had produced anemic and toothless policy proposals that focused primarily on technological solutions).

The process was detailed and litigious. The U.S. and Saudi Arabia objected to the word “confidence” in the IPCC’s proposed prediction on future warming.744 The two nations also proposed replacing a call for understanding the costs of climate change with a call for understanding its “costs and benefits,” and Saudi Arabia almost comically suggested replacing “carbon dioxide”—an inevitable product of their biggest export, oil—with the more general term “greenhouse gases.”745 OPEC countries argued that the final documents should explicitly clarify that the international community only supported

“safe” nuclear energy solutions, thereby undermining the viability of a major competitor for oil.746 India and other developing countries pressed for language pinning

responsibility for CO2-induced warming on the industrialized world, thereby

744 Leggett, The Carbon War, 17. 745 Ibid. 746 Hecht and Tirpak, “Framework Agreement,” 386-387.

317 preemptively diminishing their own responsibility for future solutions under the

UNFCCC.747 Perhaps most famously, at a meeting in Geneva in November of 1990,

Peter Timeon of the Micronesian Republic of Kiribati urged that the group endorse the idea that the greenhouse effect and sea-level rise not only “could,” but in fact “would” threaten the survival of island nations—a proposal for changing a single letter in the report that American representative John Knauss rejected, and one that nearly scuttled the final plenary session of the IPCC in the final hour.748 In each case, the debate focused on the scientific accuracy of proposed changes, but clearly the science stood proxy for a set of political and economic interests potentially at stake.

Conclusion

The awkward marriage of science and politics under the IPCC underscored a dilemma at the heart of the fight against global warming. Scientists and environmentalists continued to believe that in the right political forum, good science would inevitably lead to political change as rational political actors responded to the best information available. Successful efforts to translate international scientific consensus into international policy action on acid rain and ozone depletion reinforced this optimistic

747 Ibid. In 1990, Anil Agarwal and Sunita Narain articulated a similar sentiment that has bourgeoned into the “climate justice” movement. Writing in response to a World Resources Institute that highlighted the roles of China, India, and Brazil in contributing to global greenhouse gas emissions, Agarwal and Narain contended that ascribing responsibility to third world countries was tantamount to “environmental colonialism” wherein Western nations, reaping the benefit of years of fossil fuel use and the technologies that it has produced, blame developing countries’ deforestation, methane production, land use practices, and energy production for a significant chunk of the current and future climate crisis. Anil Agarwal and Sunita Narain, Global Warming in an Unequal World: A Case of Environmental Colonialism (New Delhi, India: Centre for Science and Environment, 1990). An excerpt of the work appears in Ken Conca, Michael Alberty, and Geoffrey D. Dabelko, Green Planet Blues: Environmental Politics from Stockholm to Rio (College Park, MD: Westview Press, 1995), 150-153. 748 For a detailed analysis of the negotiation process, see D. Bodansky, “The U.N. Framework Convention on Climate Change: A Commentary,” Yale Journal of International Law, Vol. 18, No. 2 (1993), 451-558. See also Leggett, 25-28.

318 notion. The high stakes energy and land-use issues associated with potential solutions to the problem of climate change, however, exposed the political limitations of NGO and

United Nations agency efforts to use international scientific consensus as a tool for promoting international regulations. Scientists realized that they needed strong political backing in order to implement political change on global warming. In order to establish a buy-in from the powerful state officials and government agencies that could give their recommendations real political clout, environmental advocates sought to involve these political actors in -making process. And to be fair, the effort helped to launch international political negotiations on the development of a legal framework for mitigating climate change. But the incorporation of political interests into a mechanism for creating scientific consensus also mired the scientific process in a complex web of national and international politics. Scientists soon found that their efforts to win political influence by inviting diplomats and government officials to participate in the knowledge- making process eroded their control over the knowledge-making process itself.

Scientists and environmentalists alike lauded completion of the IPCC first assessment as a transformative moment in international negotiations on climate change.

And it was an important feat. Just as UNEP and the WMO had intended, scientists and politicians had created a rigorous, transparent, and credible process for producing international scientific consensus on climate change that brought the issue international attention.749 The international community generally recognized the IPCC as the first step in developing a United Nations Framework Convention on Climate Change, to be introduced at the 1992 UNCED at Rio, and late in 1990, the United Nations General

749 See Bocking, Nature’s Experts, 124; Speth, Red Sky at Morning, 170.

319 Assembly accepted the first assessment report and officially commissioned the

Intergovernmental Negotiating Committee to begin the process of translating the IPCC’s scientific consensus into international policy. The Working Group meetings, though technically not loci of new research, helped to push new developments in atmospheric modeling, and the conversations of Working Group 1 in particular helped to expand

discussions of greenhouse gases beyond CO2 to include non-carbon gases like methane and ozone, in addition to other trace substances, that could impact the global climate system. On the whole, it is difficult to judge the IPCC fairly without recognizing its remarkable success.

At the same time, however, incorporating governments’ interests and government and industry representatives into the international scientific consensus-making process had important and not altogether positive consequences for both the science and politics of climate change. The politically-negotiated nature of the IPCC’s consensus cut two ways. On one hand, participation in the consensus-making process gave national governments the type of political ownership over the issue that Bolin and his colleagues identified as essential in any long-term climate convention process. But on the other hand, for many scientists, national-level politicians, and even some IPCC participants, the overtly political nature the IPCC tended to undermine the perceived “purity” of its consensus assessment. As Stephen Bocking describes in Nature’s Experts: Science,

Politics, and the Environment, “boundary organizations” like the IPCC are meant to

“internalize within themselves the ambiguous border between science and politics,

moderating the tendency of conflicting interests to dismantle the scientific claims of

their opponents…in effect stabilizing this border so that scientific claims are better

320 able to meet the criteria of both political and scientific credibility.”750

But in practice, scientists and politicians on all sides of the issue understood the negotiated nature of the IPCC process, and the consensus it produced carried only limited credibility within either the domestic scientific or the domestic political community (the

IPCC would later integrate a peer review process into its structure to protect the institution’s scientific credibility, to limited effect). While opponents of action on climate change had recourse to lobbying and direct participation in the international scientific consensus-making process through the IPCC, once the IPCC released its assessment, skeptics from national governments and from the energy industry also had the option to simply deny the credibility of the institution itself. 751

Secondly, though many in the media and the public perceived the IPCC as a liberal organization making bold predictions on climate change, the IPCC’s consensus-making process actually privileged a relatively conservative scientific position.752 Scientific consensus requires scientists from a given field to agree on a credible range of ideas or results that represent the state of knowledge on a given scientific problem. In an ideal world, groups of scientists would discard high and low end outliers equally. In reality, however, scientists face a greater professional risk in overstating a case than in understating it, especially when the results imply political action. The size of the IPCC,

750 Bocking, Nature’s Experts, 174. For more on boundary organizations, see also Guston, “Boundary Organizations”; van der Sluijs et al., “Anchoring Devices”; Shackley and Wynne, “Representing Uncertainty”; Sheila Jasanoff, “Contested Boundaries in Policy-Relevant Science,” Social Studies of Science, vol. 17, no. 2 (May, 1987), 195-230; Jasanoff, The Fifth Branch: Science Advisors as Policymakers (Cambridge, MA: Harvard University Press, 1990). 751 See Oreskes and Conway, Merchants of Doubt, 205-209. 752 As Oreskes and Conway write, Bert Bolin actually made a certain conservatism of language a matter of IPCC policy in order to avoid alarmist language that might lead people to discount the body’s results. Merchants of Doubt, 206. For more on the idea of “conservatism” in scientific consensus, see Robert Proctor, Cancer Wars: How Politics Shaped What We Know and Don’t Know About Cancer (New York: Basic Books, 1995), 261-265.

321 the high stakes of the economic and political implications of its reports, and the multi- tiered structure of its negotiations in the Working Groups and the plenary only exacerbated this trend toward conservative results in consensus-making.753 So too did the well-organized efforts of industry lobbyists and conservative governments water down the group’s conclusions. As a result, the first assessment report, though certainly an implicit call for policy action by its very existence, fell short of the bold recommendations international scientific leaders like Bert Bolin had hoped for.

Finally, and perhaps most importantly, the IPCC allowed political actors to continue to define the problem of climate change in terms of competing scientific views rather than the competing political and economic interests really at stake. UNEP and the

WMO did not create the IPCC as a policy mechanism; they created it as a mechanism for producing a certain form of politically-negotiated consensus knowledge. But because the

IPCC overtly presaged the creation of the UNFCCC, delegates had the opportunity to use this knowledge-making process to launch preemptive forays into the controversial political negotiations of the future. The USSR, for example—most likely one of the few

“winners” in moderate scenarios of climate change—grew increasingly uncomfortable with the limitations a strong convention on climate change would impose on the development of its oil fields. Looking for wiggle room in future negotiations, the Soviets attacked the Working Group 1 consensus or failing to include Soviet scientists’ paleoclimate data. The Soviets were not alone in their machinations. In Working Group

3, American officials attempted to shift the focus of the potential policy response in order

753 John Lanchbery and David Victor, “The Role of Science in the Global Climate Negotiations,” in Helge Ole Bergesen and Georg Parmann (eds), Green Globe Yearbook of International Co-operation and Development (Oxford: Oxford University Press, 1995), 29-40; van der Sluijs et al., “Anchoring Devices,” 314.

322 to include not just CO2, but also a wide variety of GHGs the U.S. had already begun to regulate in other contexts.754 Delegates argued over sources and sinks, gas concentrations, and atmospheric residence times, but ultimately these debates were only important to the extent that they helped determine nations’ liability to a future convention on climate change. The IPCC thus codified the consensus-making process as a de facto forum for policy debates, couched in the technical language of uncertainty and climatic complexity.

754 White House Counselor C. Gordon Bray and Assistant Attorney General Richard Stewart outlined what they called a “comprehensive approach” to climate change that included other GHGs in a Saturday seminar in Washington, D.C. during the Working Group 3 meetings. The comprehensive approach by definition included all sources and sinks of greenhouse gases, which would benefit the U.S. in two ways: first, by including the vast forested acreage on American soil, the U.S. would get credit for enormous biological sinks that required no effort on its part; and second, the comprehensive approach included gases that deplete ozone, which the U.S. had already agreed to regulate as part of the Montreal Protocol. Discussion paper: “A Comprehensive Approach to Addressing Potential Global Climate Change,” presented at an informal seminar 3 February 1990, U.S. Department of State. See also B. A. Ackerman and R. B. Stewart, “Reforming Environmental Law: The Democratic Case for Market Incentives,” Columbia Journal of Environmental Law, Vol. 13 (1988), 153-169. Both cited in Hecht and Tirpak, “Framework Agreement,” 386, 401.

323 Conclusion

The political history of global warming is particularly unsatisfying, in part because the narrative is incomplete. In the early 1990s the IPCC and the UNFCCC launched a whole new round of national and international political negotiations on global warming, and since then debate over the basic facts of global warming and about what to do about them has continued to grow and change. I have used narrative to explore the structures of advocacy on global warming as they emerged in the second half of the 20th century, and I contend that the most important intellectual, institutional, legal, and political frameworks of climate science and global warming politics were in place by

1992. But I also believe that the meat of the political history of global warming—the details that give the contemporary issue its texture and complexity—exists in a sequel to my story that has unfolded in the last two decades. The intellectual, institutional, and political structures that I describe in this dissertation have shaped this recent history, but in so doing the structures themselves have also changed. My dissertation ends in 1992, but the story of global warming politics as we know it today was only just beginning.

Even within the time period that I have chosen to investigate, I have left a few of the most difficult and important problems in the politics of global warming unexplored.

The end of the Cold War, the polarization of American politics in the 1990s, and the explosion of public, political, and academic interest in global warming have added new and important wrinkles to old problems. I take these developments as given, in part because other scholars have begun to explore these issues, but also because to treat them in due detail would draw this dissertation away from its central questions and give my narrative an episodic, Tristam Shandy sort of character.

324 Still, there is no underestimating the impacts of the profound changes in the landscapes of American and international politics in the 1990s on the politics of global warming. The most obvious and perhaps the most important of these is the end of the

Cold War. Climate science was born of the Cold War, and the politics of global warming emerged against a national and international political backdrop dominated by the conflict.

In the 1950s, atmospheric scientists defined their field in terms of a research system constructed to help fight the Cold War, and they designed their institutions at once to capitalize on that system and to subvert some of its dominant features. They shared with environmentalists a precautionary ethos and concern over unintended consequences that had deep Cold War roots. Climate scientists introduced climate change as an environmental issue in the context of a global environmental movement deeply entwined with the international politics of non-alignment and nuclear disarmament (both Cold War discourses). And finally, as scientists, environmentalists, and Congressional Democrats came together to remake the politics of global warming as a politics of dissent in the

1980s, one of their primary targets was their political opponents’ irresponsible prosecution of the same conflict, the Cold War.

In the two years between the IPCC first assessment report and the introduction of the UNFCCC, however, the dominant international political paradigm of the 20th century unexpectedly and quite suddenly vanished.755 It was replaced not by the cooperative

“new globalism” that Maurice Strong hoped to build through a strong, environmentally

755 There is a wealth of scholarship on the end of the Cold War—too many to list here—but there are a few easily accessible books that stand out within that literature. The most interesting and engaging of these, I believe, is Stephen Kotkin’s analysis of the relatively peaceful fall of the Soviet Union, Armageddon Averted: The Soviet Collapse, 1970-2000 (Oxford: Oxford University Press, 2008). For a challenging and more detailed (and clearly anti-communist) account of the dissolution of East Germany, see also Charles S. Maier’s Dissolution: The Crisis of Communism and the End of East Germany (Princeton: Princeton University Press, 1997).

325 focused United Nations, but instead by a competitive, often nationalistic, and more than anything economically oriented process of “globalization” that has reshaped both global politics and the global economy.756 Over the past twenty years, global warming advocates have been forced to navigate a rapidly changing international political landscape in which the political assumptions that helped to define their science and their advocacy simply no longer apply. It is no coincidence that Jeremy Leggett begins his book on the politics of global warming with an account of the fall of the Berlin Wall.757

Leggett uses the end of the Cold War to make an optimistic point: if a political paradigm so dominant and longstanding can collapse so quickly and unexpectedly, so too might the age of fossil fuel energy be rapidly and even peacefully brought to a close.758 I am not so sanguine. Without the ideological structures of the Cold War, global economic relationships dominate the landscape of international politics almost exclusively, and these relationships hinge upon the availability and price of the one commodity no nation can do without: energy. Upon closer investigation, I suspect that the consequences of the end of the Cold War on the fight against global warming will appear messier and more far-reaching than activists like Leggett could ever have imagined.

The end of the Cold War coincided with the of a new focus on markets and economic development within both American and international environmentalism, and this, too, has complicated the political history of global warming that I have

756 For a series of critical essays on the relationships between globalization and international environmental protection, a good start is James Gustave Speth (ed), Worlds Apart: Globalization and the Environment (Washington, D.C: Island Press, 2003). 757 Jeremy Leggett, The Carbon War: Global Warming and the End of the Oil Era (New York: Routledge, 2001), 1. Renee Panjabi also begins the first chapter of her account of the Rio Earth Summit with a discussion of the end of the Cold War as an opportunity to recontextualize international environmental affairs. Given the political history of the global environment, however, her lament that national self- interest trumped idealism at the conference seems a bit naïve. 758 Leggett, The Carbon War, 1-2.

326 described in this dissertation. In the mid-1980s, self-described “third wave” environmentalists began to look to the mechanisms of the global economy as potential tools for fighting environmental degradation. The new approach brought together a commitment to free market economic principles, a reliance on NGOs as political actors, and a deep-seated belief in the beneficence of science and technology in what Kirkpatrick

Sale calls a “systematic attempt to work with the movement’s traditional enemies, corporate polluters and extractors, to achieve by cooperation and reason what couldn’t be done by confrontation and regulation.”759 In the U.S., third wave environmentalists coupled their local and national lobbying and litigation efforts with a new campaign to engage directly with the business community to demonstrate the economic benefits of more efficient, environmentally responsible business practices. Abroad, they brought a similar economic focus to the dual problems of environmental degradation and economic poverty under the rubric of “sustainable development,” a catch-all term for strategies to promote both economic and environmental well-being through more efficient resource allocation.760 When the United Nations introduced the UNFCCC to the world in 1992, they did so in the context of the first major international meeting of the post-Cold War era: the U.N. Conference on Environment and Development. It was a conference built around the sustainable development concept.

759 Kirkpatrick Sale, The Green Revolution: The American Environmental Movement, 1962-1992 (New York: Hill & Wang, 1993), 83. See also Samuel P. Hays’ Beauty, Health, and Permanence: Environmental Politics in the United States, 1955-1985 (New York: Cambridge University Press, 1987), 512. 760 More than any other document, the Brundtland Report laid the foundations of the sustainable development concept in international politics. “Humanity has the ability to make development sustainable—to ensure that it meets the needs of the present without compromising the ability of future generations to meet their own needs,” the WCED wrote. “People can build a future that is more prosperous, more just, and more secure,” ushering in “a new era of economic growth…based on policies that sustain and expand the environmental resource base.” World Commission on Environment and Development, Our Common Future (Oxford: Oxford University Press, 1987), ES-7, ES-1, 4.

327 The consequences of the rise of market environmentalism for the politics of global warming, like those of the end of the Cold War, are murky, and both subjects beg for more research and for more hard thinking. On the surface, combining the goals of economic development and environmental protection privileged the interests of the Third

World, and in that sense the Rio Earth Summit helped to address the sort of environmental justice and equity issues that had nearly derailed the 1972 U.N.

Conference on the Human Environment. Third wave environmentalists had already begun to find success in harnessing the economic mechanisms of market to increase energy efficiency, to reduce waste, and even to protect species and habitats in a number of places.761 Their work with the World Bank and with other private national and international financial institutions to make capital available for projects mutually beneficial to economies and environments helped to give Third World nations a stake in environmental affairs.

But “sustainable development” is a concept closely tied with the promulgation of a high-consumption First World standard of living.762 Even if we can ameliorate the

761 See Sale, The Green Revolution, 85; Dana R. Visser and Guillermo A. Mendoza, “Debt-for-Nature Swaps in Latin America,” Journal of Forestry, vol. 92, no. 6 (November, 1994): 13-16; Robert T. Deacon and Paul Murphy, “The Structure of an Environmental Transaction: The Debt-for-Nature Swap,” Land Economics vol. 73, no. 1 (February, 1997):1-24. 762 In the 1980s, so-called “deep ecologists” articulated a convincing critique of sustainable development, much of which still applies today. See Arne Naess, “Sustainable Development and the Deep Ecology Movement,” in Sustainable Development: Theory, Policy, and Practice Within the European Union, edited by Susan Baker, Maria Kousis, Dick Richardson, and Steven Young (London: Routledge, 1997), 61-71; Bill Devall and George Sessions (eds), Deep Ecology: Living as if Nature Mattered (Salt Lake City: Peregrine Smith Books, 1985); Arne Naess, “Sustainable Development and Deep Ecology,” in Ethics of Environment and Development: Global Challenge, International Response, edited by J. Ronald Engel and Joan Gibb Engel (London: John Wiley, 1992). For a very basic overview of deep ecology, see also “Arne Naess on Deep Ecology, 1982, 1984,” in The Environmental Debate: A Documentary History, edited by Peninah Neimark and Peter Rhoades Mott (Westport, CT: Greenwood Press, 1999): 247-249. Deep Ecology itself was a hotly contested idea throughout the 1980s; the concept received criticism from the mainstream environmental movement from being politically untenable, from third-wavers for being ineffective, and from social ecologists, most notably Murray Bookchin, for ignoring the centrality of social paradigms—capitalism, authoritarianism, hierarchy—to ecological behavior. Bookchin is probably the

328 environmental impacts of the stages of future development, bringing the Third World

“up” to First World living conditions eventually requires vast material resources and, more than anything else, energy. Moreover, framing the fight against global warming in any kind of a development framework foists responsibility for the problem upon the developing world.763 In reality, it is not only a developing world problem; it is a global problem. Or, to put a finer point on it, it is a problem with global consequences—uneven consequences stemming largely from the economic success of the First World over time.

While curbing CO2 emissions in the future requires an international political effort to reduce the climatic impact of development in places like India and China, that does not absolve the long-ago industrialized nations of North America and Europe from their responsibilities for our current crisis. The division between the haves of the industrialized world and the have nots of the developing world is an old one in international environmental affairs. In the absence of the Cold War, it has also now become the dominant feature of the international political landscape of global warming.764

If the political landscape of global warming has changed, however, many of the key structures that characterized its emergence as a scientific, political, and

most famous critic of deep ecology, in part because of his 1989 debate with Dave Foreman on the subject. See David Levine (ed), Defending the Earth: A Dialogue Between Murray Bookchin and Dave Foreman (Boston, MA: South End Press, 1991); Murray Bookchin, “Social Ecology vs. Deep Ecology: A Challenge for the Ecology Movement,” Green Perspectives: Newsletter of the Green Program Project, nos. 4-5 (summer 1987). 763 The concept of “climate justice” was articulated relatively early in the development of the IPCC- UNFCCC regime of international climate law in Anil Agarwal and Sunita Narain, Global Warming in an Unequal World: A Case of Environmental Colonialism (New Delhi, India: Centre for Science and Environment, 1990). An excerpt of the work appears in Ken Conca, Michael Alberty, and Geoffrey D. Dabelko, Green Planet Blues: Environmental Politics from Stockholm to Rio (College Park, MD: Westview Press, 1995), 150-153. 764 See Ranee K.L. Panjabi, “The South and the Earth Summit: The Development/Environment Dichotomy,” in The Earth Summit at Rio: Politics, Economics, and the Environment (Boston: Northeastern University Press, 1997), 93-155; James Gustave Speth and Peter M. Haas, Global Environmental Governance (Washington, D.C.: Island Press, 2006); Speth, Worlds Apart; Agarwal and Narain, Global Warming in an Unequal World.

329 environmental issue in the 20th century remain. The first and most important structural constant in the politics of climate change is the primacy of science. Climate science continues to serve as a surrogate for climate politics largely because the economic, legal, and scientific mechanisms that we have developed to deal with the global environment—and with global warming more specifically—all revolve around the production and validation of knowledge. In the 1980s and early ‘90s, third wave environmentalists mobilized banks and markets to protect forests and reduce waste, but their primary response to climate change was to set up thinktanks and to produce not policy or political change but information. National and international environmental

NGOs advocated a framework convention on climate change, but nearly all parties consistently identified “fact-finding” and consensus-making as the necessary first steps of any legal or regulatory regime. Environmentalists and politicians looked to scientists for answers, and scientists, being scientists, responded by devising new and more politically convincing ways to produce and present consensus knowledge. In the absence of strong and specific policies and regulations to control energy production, land use, and emissions, mechanisms of knowledge production like the IPCC served as not just the primary forum for international negotiations on climate change, but really the only forum for those discussions.

Even in the two decades since advent of the UNFCCC—an international legal framework designed specifically to guide political actors from science to international policy—the politics of global warming has continued to revolve around knowledge and consensus. Indeed, the UNFCCC’s authors integrated the primacy of science into the structure of the treaty itself. Six out of the ten “commitments” of Article Four of the

330 convention emphasized informational and educational goals. 765 Articles Five, Six, Nine, and Twelve also dealt specifically with creating and distributing knowledge.766 Though the treaty maintained a clear overall objective—the “stabilization of greenhouse gas concentrations in the atmosphere at a level that would prevent dangerous anthropogenic interference with the climate system”—the UNFCCC’s non-scientific commitments remained relatively vague. Any implementation of these commitments relied on protocols—specifically the Kyoto Protocol—to provide specific numerical targets for greenhouse gas emissions, again based on the most up-to-date, politically negotiated science. The process required a renewed agreement on the nature, causes, risks, and possible solutions to climate change in advance of the national-level ratification process, and again, nations party to the UNFCCC turned to the IPCC—that is, to science—for answers.767

The centrality of science and scientists in global warming politics over the last fifty years has had a number of consequences. It has forced scientists to play unfamiliar roles as political advocates and to develop unfamiliar hybrid institutions like the IPCC that overtly and intentionally blend science with politics. Scientists’ unique access to the global atmosphere has inspired debates within the climate science community over sticky concepts about “good science” and the limitations and responsibilities of politically

765 “Annex: United Nations Framework Convention on Climate Change,” in Sten Nilsson and David Pitt, Protecting the Atmosphere: The Climate Change Convention and its Context (London: Earthscan Publications, 1994), 177-203. 766 Ibid. 767 The ongoing relationship between the IPCC and UNFCCC only reinforced the centrality of science in climate politics in the 1990s. Scientists continued to compile and evaluate evidence on climate change in order to negotiate a dynamic and ongoing form of consensus in parallel to the Conferences of the Parties (COPs) of the UNFCCC. Changes in the IPCC review process have facilitated smoother day-to-day operations of this mechanism, but even today scientists and political delegates continue to replay the political debates and negotiations of the first assessment report with each new scientific and political development on climate change. Annex: United Nations Framework Convention on Climate Change,” in Nilsson and Pitt, Protecting the Atmosphere, 181.

331 active scientific professionals. Scientists’ engagement in politics, and especially in international politics via the IPCC, has changed the character and meaning of scientific consensus. Global warming advocates’ continued reliance on the power of good science has left them vulnerable to attack from political conservatives, energy-intensive industries, and corporations, as well as from other scientists. Often in this history, the politics that climate science has been made to support has been difficult to meaningfully distinguish from the science itself.

The highly technical, scientific nature of global warming has complicated the character of American environmentalism, too. It is no longer the primarily grassroots consumer movement focused on protecting “amenities” like clean air, clean water, open space, and access to wilderness and the wilderness experience that Sam Hays described in

1987.768 It has become more global, more market-oriented, and more scientific.

Considering their slow and halting engagement with global warming in the first half of the 20th century, it is remarkable the extent to which American environmental organizations have embraced the issue in the last twenty years. The Sierra Club, for example, was in 1980 not at all interested in tackling such a sticky global issue, but in the first decade of the 21st century it has made climate change its top priority.769 Few environmental organizations have failed to address climate change as a programmatic focus, and the issue now drives the broader movement more than any other. For better or worse, environmentalists too now trade in the language of science—of ppm (parts per million), GHGe (greenhouse gas equivalent), BTU (British Thermal Units), and climate

768 Hays, Beauty, Health, and Permanence. 769 A description of the Sierra Club’s “Climate Recovery Partnership” can be found at http://www.sierraclub.org/crp/campaigns/default.aspx. See Kate Galbraith, “An Environmentalist Steers Towards the Middle,” The New York Times, October 15, 2008; Kate Galbraith, “Carl Pope Switches Roles at the Sierra Club,” The New York Times, January 23, 2009.

332 forcings—as much as they do in concepts like nature and wilderness that have historically shaped the movement’s grassroots campaigns and core values.

Finally, because scientists more than any other group have shaped the politics of global warming, the scale at which scientists understand the problem has shaped the scale of the political response to the issue. As Stephen Bocking argues, the scientists, environmentalists, and politicians involved in developing the IPCC and UNFCCC all assumed that science “provided not just a description of the problem, but a template for action.”770 As a physical phenomenon and subject of study, the spatial and temporal scales of climate change have over the last half-century forced scientists to establish new global networks of environmental monitoring, to build simulations of the global atmosphere, and to establish better cooperation among national and international scientific organizations. When scientists described the global nature of the problem—including its global causes and its global impacts—they took for granted that the ensuing political solutions must also be global.771

But the term “global” is as problematic as is now iconic in the fight against global warming. Real political solutions do not exist at a truly global scale; rather, “global” political solutions represent agreements between individual nations and groups of nations to behave in certain ways on the national level. Constructed at the scale that scientists understand climate change, the UNFCCC and Kyoto Protocol serve to coordinate national efforts to reduce emissions and to provide goals for those reductions.

Ultimately, the prosecution of those reductions—that is, real action on climate

770 Bocking, Nature’s Experts, 132. 771 Lawerence H. Goulder and Brian M. Nadreau, “International Approaches to Reducing Greenhouse Gas Emissions,” in Climate Change Policy: A Survey, edited by Stephen H. Schneider, Armin Rosencranz, and John O. Niles (Washington, D.C.: Island Press, 2002), pg. 115.

333 change—remains the responsibility of nations, states, and communities. The problem is global; its solutions, to the extent that they exist, must be local.

The disconnect between the global nature of the problem and individuals’ typically local experience of the world stands at the crux of the political problem that I have described in this dissertation. As Dipesh Chakrabarty argues in “The Climate of

History: Four Theses,” humans are historically unaccustomed to thinking about themselves as part of a collective global species, let alone a species that has attained agency on a global scale. Both of these perceptions—that of the self as part of a species and that of the species as a geological agent—defy the individuality and geological powerlessness that have characterized individual experiences for the vast majority of human history. Climate change challenges us to accept both of these unfamiliar concepts. For Chakrabarty, this means approaching not only environmental politics, but human experience itself, in a different way.772

A noble idea, but not, I argue, a politically tenable one. In his analysis of 20th century American environmentalism, Hays identifies the city, wildlands, and the countryside as “important organizing principles for examining the political sociology of environmental ideas and action,” and essentially, these three types of spaces create a framework for understanding regional environmental issues throughout the 20th and early

21st centuries.773 As Chakrabarty demonstrates, atmospheric change continues to defy these categories of direct environmental experience, if for no other reason than it defies their geographical boundaries. But these categories of geographical organization are useful—in fact essential—in environmental politics, because they help to tie political

772 Dipesh Chakrabarty, “Climate and History: Four Theses,” Critical Inquiry 35 (Winter 2009), pg. 206. 773 Hays, Beauty, Health, and Permanence, 8.

334 constituents to geographically-specific environmental problems. They are constituents at once attached to the environment and to the social and political structures of communities, towns, cities, and even regions. A the national level, environmental organizations working with and against the bureaucracy have become politically effective and politically powerful, but at its core, these links between political constituents and discrete geographical spaces are what has historically given the American environmental movement its muscle. As both the 1972 U.N. Conference on the Human Environment and the 1992 Rio Earth Summit have shown, so, too, have geographically discrete issues driven international environmentalism.774 So while it is heartening that environmental organizations like the Sierra Club and Wilderness Society have moved to take action on global warming, I find it alarming that they seem to have sacrificed their local, value- oriented approach to do so.

In fact, a return to a local, community-based form of environmentalism may be exactly the kind of politics that are needed to complement the push toward global

solutions to climate change. While the climate system is global, and while CO2-induced warming affects the Earth through global-scale geophysical processes, the root cause of global warming—the burning of fossil fuels and the emission of other GHGs—is also inherently local. Energy use around the world over time has contributed to a global

increase in atmospheric CO2, and thus humans in the aggregate—“historically and collectively,” as Dipesh Chakrabarty puts it—cause global warming.775 But people do not use energy globally and over the long term; they use it locally and immediately, to produce heat and light, to manufacture goods, and for transportation, all more or less at

774 See also Christof Mauch, Nathan Stoltzfus, and Douglas R. Weiner (eds), Shades of Green: Environmental Activism Around the Globe (Lanham, MD: Rowman & Littlefield Publishers, 2006). 775 Chakrabarty, “The Climate of History,” 206.

335 the scale of individual human experience. They consume products within a world economy derived from long, often multinational manufacturing processes requiring energy for both their making and for their transport, but again, they consume these products in discrete times and places—that is, locally.

That is not to say that we should stop working toward global solutions, nor that scientists should cease to play a primary role in building these solutions. On the contrary, institutions like the IPCC and treaties like the UNFCCC provide both guidance and a political justification for the national-level efforts essential to curbing emissions worldwide. Science helps to give these international agreements their authority, and at this macro level, science and its numerical language continue to help define the problem and measure our collective progress—or lack of progress—in solving it.

But information is not the same as political change. Scientists, environmentalists, and international leaders have helped to create and have participated in international organizations designed to make and apply knowledge, but these institutions lack the political power to meaningfully affect the production and consumption of energy at its most important local, regional, and national scales. That power remains, not surprisingly, local, regional, and national, and at those scales, leaders have been reluctant to follow scientists’ and international leaders’ quasi-legal recommendations.

Ultimately, science and international political agreements alone cannot provide solutions to the problem because the problem is not exclusively or even primarily scientific. I have focused on the structures of science and political advocacy in this dissertation because those have been the dominant forces in global warming discourse for the last half-century, but at its core the problem is at once cultural, political, scientific and

336 economic. So far, we have mostly tinkered—tinkered with technology, tinkered with discrete components of our energy systems and political institutions, tinkered with the science. And to be fair, small changes in things like coal technology, increases in automobile fuel efficiency, new renewable energy sources, and more convincing scientific institutions make incremental differences, and they are both practically and symbolically important.

But the problem is not the tinkering kind. Global warming is so alarming because real solutions involve much more fundamental social and cultural change. It is a problem tied to what we eat and where we eat it. It involves questions about how and why we move around and about the things that we produce and how and where we consume them.

A single house, or car, or lawn, or farm may have little impact on global CO2 as a whole, but global warming it is very much a question about our houses, our cars, our lawns, and our farms. It is about what we value and why we value it. To an extent, these are still questions of science and economics. We need to be able to weigh our current cultural practices against the risks they pose to our values. But they are also difficult questions of politics, of ethics, and of morality. Scientists studying climate have tended to ignore these questions, or to mask them in the seemingly neutral language of numbers. But it is the complex answers to these local human questions that shape our relationship to the

Earth and its global systems. It is these value-laden questions that also must shape our local, national, and global responses to global warming as we move forward.

337 Bibliography

Archival Collections AAAS Climate Program Records, American Association for the Advancement of Science Archives, Washington, D.C. AAAS Committee on New Directions Files, American Association for the Advancement of Science Archives, Washington, D.C. George A. Keyworth, III Papers, Ronald Reagan Archives, Ronald Reagan Presidential Library, Simi Valley, CA James A. Baker, III Files, Ronald Reagan Archives, Ronald Reagan Presidential Library, Simi Valley, CA. John C. Whitaker Papers, Nixon Presidential Materials Project, White House Central Files, Staff Member Office Files, College Park. MD. John Firor Papers, UCAR/NCAR Archives, Boulder, CO. NCAR Nuclear Winter Files, UCAR/NCAR Archives, Boulder, CO. Peter Galbraith Working Files, Committee on Foreign Relations, Records of the U.S. Senate, National Archives and Records Administration (NARA-1), Washington, D.C. Philip D. Thompson Papers, UCAR/NCAR Archives, Boulder, CO. Roger Randall Dougan Revelle Papers, Scripps Institution of Oceanography Archives, La Jolla, CA. Sierra Club General Records, Bancroft Library, University of California, Berkeley, CA. Sierra Club International Records, Bancroft Library, University of California, Berkeley, CA. Stephen Schneider Papers, UCAR/NCAR Archives, Boulder, CO. Stephen Schneider Personal Files, by permission, Stanford, CA. William W. Kellogg Files, UCAR/NCAR Archives, Boulder, CO. Walter Orr Roberts FBI File (FOIA), UCAR/NCAR Archives, Boulder, CO. Walter Orr Roberts Papers, University of Colorado Archives, Bounder, CO.

Selected Interviews Arrhenius, Gustaf Olof. “Oral History of Gustaf Olof Arrhenius.” By Laura Harkewicz. April 11, 2006. Scripps Institute of Oceanography Archives. http://libraries.ucsd.edu/locations/sio/scripps- archives/resources/collections/oral.html

Broecker, Wallace. By Spencer Weart at the Lamont-Doherty Earth Observatory, NY. November 14, 1997. Niels Bohr Library & Archives, American Institute of Physics, College Park, MD USA. http://www.aip.org/history/ohilist/23909_1.html.

Firor, John. By Earl Droessler. June 26, 1990. UCAR/NCAR Oral History Project.

338 Glantz, Michael. November 12, 2007. National Center for Atmospheric Research, Boulder, CO. With the author. –––. April 14, 2008. With the author, by telephone.

Kellogg, William. By Ed Wolff and Nancy Gauss. February 10, 1980. NCAR.UCAR Oral History Project

Hansen, Jim. By Spencer Weart at Goddard Institute for Space Studies, New York City. October 23, 2000. Niels Bohr Library & Archives, American Institute of Physics, College Park, MD USA, http://www.aip.org/history/ohilist/24309_1.html

Hayes, Denis. June 2, 2009. Seattle, WA. With the author.

Machta, Lester. By Julius London. October 31, 1993. AMS/UCAR Tape Recorded Interview Project.

Malone, Thomas F. By Earl Droessler. February 18, 1989. AMS/UCAR Tape Recorded Interview Project. –––. June 4, 2009. With the author, by telephone.

Revelle, Roger R. “Preparation for a Scientific Career.” Oral history by Sarah Sharp. 1984. Regional Oral History Office, The Bancroft Library, University of California, Berkeley, 1988. S10 Reference Series No. 88, November, 1988.

Revelle, Roger. By Earl Droessler. February, 1989. American Institute of Physics, College Park, MD.

Roberts, Walter Orr. “Early History of the High Altitude Observatory.” July 5, 1966. UCAR oral History Project. UCAR/NCAR Archives. –––. By David H. DeVorkin. July 26-28, 1983. Oral History Program, National Air and Space Museum of the Smithsonian Institution.

Robinson, George D. By Earl Droessler. June 27-28, 1994. AMS/UCAR Tape Recorded Interview Project.

Schneider, Stephen. By Bob Chervin. January 10-13, 2002. AMS/UCAR Tape Recorded Interview Project.

Thomson, Phil and Ed Lorenz. “Dialogue between Phil Thompson and Ed Lorenz.” July 31, 1986. AMS/UCAR Tape Recorded Interview Project.

Smagorinski, Joseph. By John Young. May 16, 1986. AMS/UCAR Tape Recorded Interview Project.

339 Published Sources

Abramson, Albert. Zworykin: Pioneer of Television. Champaign, IL: University of Illinois Press, 1995.

Ackerman, B.A and R.B. Stewart. “Reforming Environmental Law: The Democratic Case for Market Incentives.” Columbia Journal of Environmental Law, vol. 13, no. 171 (1988): 153-169.

Agarwal, Anil and Sunita Narain. Global Warming in an Unequal World: A Case of Environmental Colonialism. New Delhi, India: Centre for Science and Environment, 1990.

Agrawala, Shardul. “Context and Early Origins of the Intergovernmental Panel on Climate Change.” Climatic Change, vol. 39, no. 4 (August, 1998): 605-620.

Allen, Jonathan, (ed). March 4: Scientists, Students, and Society. Cambridge: MIT Press, 1970.

Allison, Graham T. and Philip Zelikow. Essence of a Decision: Explaining the Cuban Missile Crisis. 2nd ed. New York: Longman, 1999.

Alvarez, Luis W. Alvarez: Adventures of a Physicist. New York: Basic, 1987.

Alvarez, Lauis W., Walter Alvarez, Frank Asaro, and Helen V. Michel. “Extraterrestrial Causes for the Cretaceous-Tertiary Extinction.” Science , vol. 208, no. 4448 (Jun. 6, 1980): 1095-1108.

American Association for the Advancement of Science. “AAAS Resolution: Study of Ecological Effects of Herbicides.” December 30, 1968. About AAAS: History & Archives. http://archives.aaas.org/docs/resolutions.php?doc_id=279. –––. “AAAS Resolution: Use of Herbicides in Vietnam: Call for Field Study.” December 30, 1969. About AAAS: History & Archives. http://archives.aaas.org/docs/resolutions.php?doc_id=283. –––. “AAAS Resolution: Commending U.S. Government Phase-Out of Herbicides in Vietnam.” December 30, 1970. About AAAS: History & Archives. http://archives.aaas.org/docs/resolutions.php?doc_id=287. –––. “AAAS Resolution: Assessment of the Ecological Effects of U.S. Activities in Vietnam.” December 30, 1972. About AAAS: History & Archives. http://archives.aaas.org/docs/resolutions.php?doc_id=296. –––. “AAAS Resolution: Appeal for Cessation of Hostilities in Vietnam.” December 30, 1972. About AAAS: History & Archives. http://archives.aaas.org/docs/resolutions.php?doc_id=294.

340 Arrhenius, Svante. “On the Influence of Carbonic Acid in the Air Upon the Temperature of the Ground.” The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science, ser. 5 (April 1896): 237-276. –––. Worlds in the Making: The Evolution of the Universe. Translated by H. Borns. New York: Harper and Brothers, 1908.

Badash, Lawrence. Science and the Development of Nuclear Weapons: From Fission to the Limited Test Ban Treaty, 1939-1963. Atlantic Highlands, NJ: Humanities Press, 1995. –––. “Nuclear Winter: Scientists in the Political Arena,” Physics in Perspective, vol. 3, no. 1 (March, 2001): 76-105. –––. A Nuclear Winter’s Tale: Science and Politics in the 1980s. Cambridge, MA: MIT Press, 2009.

Barney, Gerald O., ed. Global 2000 Report to the President: Entering the 21st Century. Washington, D.C.: U.S. Government Printing Office, 1980.

Barrow, John C. “An Age of Limits: Jimmy Carter and the Quest for a National Energy Policy.” In The Carter Presidency: Policy Choices in the Post-New Deal Era, edited by Gary M. Fink and Hugh Davis Graham, 158-223. Lawrence, KS: University Press of Kansas, 1998.

Batchelor, G.K. An Introduction to Fluid Dynamics. Cambridge: Cambridge University Press, 1967.

Beck, Roy and Leon Kolankiewicz. “The Environmental Movement’s Retreat From Advocating U.S. Population Stabilization (1970-1998): A First Draft of History.” Journal of Policy History, vol. 12, no. 1 (2000), 123-156.

Berger, A. “Milankovitch Theory and Climate.” Review of Geophysics, vol. 26, no. 4 (1988): 624-57.

Bird, Kai and Martin J. Sherwin. American Prometheus: The Triumph and Tragedy of J. Robert Oppenheimer. New York: Random House, 2005.

Bocking, Stephen. Nature’s Experts: Science, Politics, and the Environment. New Brunswick: Rutgers University Press, 2006.

Bodansky, D. “The U.N. Framework Convention on Climate Change: A Commentary.” Yale Journal of International Law, vol. 18, no. 2 (1993): 451-558.

Bolin, Bert and Erik Eriksson. “Changes in the Carbon Dioxide Content of the Atmosphere and Sea due to Fossil Fuel Combustion.” 1958. In The Atmosphere and the Sea in Motion: Scientific Contributions to the Rossby Memorial Volume, edited by Bert Bolin, 130-142. New York: Rockefeller Institute Press, 1959. Also at http://wiki.nsdl.org/index.php/PALE:ClassicArticles/GlobalWarming/Article8.

341 Bookchin, Murray. “Social Ecology vs. Deep Ecology: A Challenge for the Ecology Movement.” Green Perspectives: Newsletter of the Green Program Project, no. 4 (Summer 1987): 1-13.

Botkin, Daniel B. Discordant Harmonies: A New Ecology for the Twenty-First Century. Oxford: Oxford University Press, 1990.

Boulding, Kenneth. “The Economics of the Coming Spaceship Earth.” In Environmental Quality in a Growing Economy, Essays from the Sixth RFF Forum, edited by Henry Jarrett, 3-14. Baltimore: Johns Hopkins Press, 1966.

Bowen, Mark. Censoring Science: Inside the Political Attack on Dr. James Hansen and the Truth of Global Warming. New York: Plume, 2008.

Boyle, Robert H. and R. Alexander Boyle. Acid Rain. New York: Schocken Books, 1983.

Brammer, H. “Floods in Bangladesh: Geographical Background to the 1987 and 1988 Floods.” The Geographical Journal, vol. 156, no. 1 (March, 1990): 12-22.

Brandt, Allan. The Cigarette Century: The Rise, Fall, and Deadly Persistence of the Product that Defined America. Boulder, CO: Basic Books, 2007.

Braudel, Fernand. The Mediterranean and the Mediterranean World in the Age of Philip II. Translated by Siân Reynolds. New York: Harper and Row, 1972.

Broecker, Wallace S. “Climatic Change: Are We on the Brink of a Pronounced Global Warming?” Science, vol. 189, no. 4201 (August 8, 1975): 460.

Broecker, Wallace S., David L. Thurber, John Goddard, Teh-Lung Ku, R. K. Matthews and Kenneth J. Mesolella. “Milankovitch Hypothesis Supported by Precise Dating of Coral Reef and Deep-Sea Sediments.” Science, vol. 159, no. 3812 (January 19, 1968): 297-300.

Brooks, C.E.P. Climate Through the Ages: A Study of the Climatic Factors and Their Variations. 2nd Revised Edition. New York: Dover Publications, 1970.

Brown, Lester. By Bread Alone. With Erik P. Eckholm. New York: Praeger, 1974.

Bryson, Reid A. and Wayne M. Wendland. “Climatic Effects of Atmospheric Pollution.” In Global Effects of Environmental Pollution, edited by S. Fred Singer, 130-138. New York: Springer-Verlag, 1970.

Bryson, Reid E. and Thomas J. Murray. Climates of Hunger: Mankind and the World’s Changing Weather. Madison: University of Wisconsin Press, 1977.

342 Bundy, McGeorge. Danger and Survival: Choices About the Bomb in the First Fifty Years. New York: Random House, 1988.

Burks, Alice Rowe. Who Invented the Computer? The Legal Battle That Changed Computing History. New York: Prometheus Books, 2003.

Calder, Nigel. The Weather Machine. New York: Viking, 1975.

Callendar, Guy Stewart. “The Artificial Production of Carbon Dioxide and Its Influence on Climate.” Quarterly Journal of the Royal Meteorological Society 64 (1938):223-240.

Carson, Rachel. The Sea Around Us. New York: New American Library, 1954. –––. Silent Spring. Boston: Houghton Mifflin, 1962.

Cattell, J. McKeen. “The Organization of Scientific Men.” The Scientific Monthly, vol. 14, no. 6 (June 1922): 568-578.

Center for Biological Diversity. “Secretary of the Interior Dirk Kempthorne Wins 2007 Rubber Dodo Award.” Press Release, August 24, 2007. http://www.biologicaldiversity.org/news/press_releases/kempthorne-08-24- 2007.html.

Chakrabarty, Dipesh. “The Climate of History: Four Theses.” Critical Inquiry 35 (Winter, 2009), 197-222.

Christiansen, Norman L., James K. Agee, Peter F. Brussard, Jay Hughes, Dennis H. Knight, G. Wayne Minshall, James M. Peek, Stephen J. Pyne, Frederick J. Swanson, Jack Ward Thomas, Stephen Wells, Stephen E. Williams and Henry A. Wright. “Interpreting the Yellowstone Fires of 1988.” Bioscience, vol. 39, no. 10 (November, 1989): 678-685.

Christianson, Gale E. Greenhouse: the 200-year Story of Global Warming. New York: Penguin Books, 1999.

Chung, Jan and Jon Halliday. Mao: The Unknown Story. New York: Random House, 2005.

Cohen, Linda R. and Roger G. Noll. The Technology Pork Barrel. Washington: The Brookings Institution, 1991.

Commoner, Barry. The Closing Circle: Nature, Man, and Technology. New York: Alfred A. Knopf, Inc., 1971.

343 Conca, Ken, Michael Alberty, and Geoffrey D. Dabelko. Green Planet Blues: Environmental Politics from Stockholm to Rio. College Park, MD: Westview Press, 1995.

Conservation Foundation. Implications of Rising Carbon Dioxide Content of the Atmosphere. New York: The Conservation Foundation, 1963.

Conway, Erik. High Speed Dreams: NASA and the Technopolitics of Supersonic Transportation, 1945-1999. Baltimore: Johns Hopkins University Press, 2005. –––. A History of Atmospheric Sciences at NASA: A History. Baltimore: Johns Hopkins Press, 2008.

Cowling, Ellis B. “Acid Precipitation in Historical Perspective.” Environmental Science and Technology, vol. 16, no. 2 (February, 1982): 111A.

Cronon, William (ed). Uncommon Ground: Toward Reinventing Nature. New York: W.W. Norton and Co., 1995.

Crutzen, P. J. “The Influence of Nitrogen Oxides on the Atmospheric Ozone Content.” Quarterly Journal of the Royal Meteorological Society, vol. 96, no. 408 (December 1970): 320–325.

Crutzen, Paul J. and John W. Birks. “The Atmosphere After a Nuclear War: Twilight at Noon.” Ambio, vol 11, no. 2/3 (Winter, 1982): 114-125.

Cullather, Nick. “Miracles of Modernization: The Green Revolution and the Apotheosis of Technology.” Diplomatic History, vol. 28, no. 2 (April 2004): 227-54.

Curtis, Carolyn, ed. Before the Rainbow: What we Know About Acid Rain. Washington: Edison Electric Institute, 1980.

Davidson, Keay. Carl Sagan: A Life. New York: Wiley, 2000.

Day, Deborah. Roger Randall Dougan Revelle Biography. UC San Diego: Scripps Institution of Oceanography Archives, 2008. http://escholarship.org/uc/item/78d9v14v

De, U.S., R.K.Dube and G.S.Prakasa Rao. “Extreme Weather Events Over India in the Last 100 Years.” Journal of the Indian Geophysical Union, vol. 9, no. 3 (July, 2005): 173-187.

Deacon, Robert T. and Paul Murphy. “The Structure of an Environmental Transaction: The Debt-for-Nature Swap.” Land Economics vol. 73, no. 1 (February, 1997):1- 24.

344 Devall, Bill and George Sessions, eds. Deep Ecology: Living as if Nature Mattered. Salt Lake City: Peregrine Smith Books, 1985.

Development and Environment: Report and Working Papers of a Panel of Experts Convened by the Secretary-General of the United Nations Conference on the Human Environment (Founex, Switzerland, June 4-12, 1971). Paris: Mouton, 1972. (also known as the “Founex Report.”)

Donahue, Thomas M. and Alan J. Grobecker. “The SST and Ozone Depletion.” Science, vol. 187, no. 4182 (March 28, 1975): 1142.

Dotto, Lydia and Harold Shriff. The Ozone War. Garden City, NY: Doubleday, 1978

Druffel, Ann and Jaques Vallee. Firestorm: Dr. James E. McDonald's Fight for UFO Science. Columbus, NC: Wild Flower Press, 2003.

Dubos, Rene and Barbara Ward. Only One Earth: The Care and Maintenance of a Small Planet. New York: W.W. Norton & Co, 1972.

Dwiggings, Don. The SST: Here it Comes Ready or Not. New York: Doubleday, 1968.

Egan, John. Barry Commoner and the Science of Survival. Cambridge: MIT Press, 2007.

Ehrlich, Paul. The Population Bomb. New York: Ballantine Books, 1968.

Ehrlich, Paul R. and Anne H. Ehrlich. Population, Resources, Environment: Issues in Human Ecology. San Francisco: W.H. Freeman and Company, 1970.

Ehrlich, Paul R. and John P. Holdren. “Dispute.” Environment 14 (April 1972): 23-52.

Ehrlich, Paul R., John Harte, Mark A. Harwell, Peter H. Raven, Carl Sagan, George M. Woodwell, Joseph Berry, et al. “Long-Term Biological Consequences of Nuclear War.” Science, vol. 222, no. 4630 (December 23, 1983): 1293-1300.

Ehrlich, Paul, Carl Sagan, Donald Kennedy, and Waler Orr Roberts. The Cold and the Dark: The World After Nuclear War, Conference on the Long-Term Worldwide Consequences of Nuclear War, Washington, D.C., 1983. New York: Norton, 1984.

Elichirigoity, Fernando. Planet Management: Limits to Growth, Computer Simulation, and the Emergence of Global Spaces. Evanston, IL: Northwestern University Press, 1999.

Ewards, Paul. “The World in a Machine.” In Systems, Experts, and Computers: The Systems Approach in Management and Engineering, World War II and After.

345 Edited by Agatha Hughes and Thomas Hughes, 221-253. Cambridge: MIT Press, 2000.

Fagan, Brian. The Long Summer: How Climate Changed Civilization. New York: Basic Books, 2004. –––. The Little Ice Age: How Climate Made History, 1300-1850. New York: Basic Books, 2000.

Farman, J.C., B.G. Gardiner, and J.D. Shanklin. “Large Losses of Total Ozone in

Antarctica Reveal Seasonal ClOx/NOx Interaction.” Nature 315 (May 16, 1985): 207-10.

Finch, Boyd L. Legacies of Camelot: Stewart and Lee Udall, American Culture, and the Arts. Norman: University of Oklahoma Press, 2008.

Fitzgerald, Frances. Way Out There in the Blue: Reagan, Star Wars, and the End of the Cold War. New York: Simon and Schuster, 2000.

Flannery, Tim. The Weather Makers: How Man is Changing the Climate and What it Means for Life on Earth. New York: Atlantic Monthly Press, 2005.

Fleming, James Roger. Historical Perspectives on Climate Change. Oxford: Oxford University Press, 1998. –––. The Callendar Effect: The Life and Times of Guy Stewart Callendar (1898-1964), the Scientist Who Established the Carbon Dioxide Theory of Climate Change. Boston: American Meteorological Society, 2007. –––. “The Climate Engineers: Playing God to Save the Planet.” The Wilson Quarterly (Spring 2007): 46-60.

Flippin, J. Brooks. Nixon and the Environment. Albuquerque: University of New Mexico Press, 2000. –––. Conservative Conservationists: Russell Train and the Emergence of American Environmentalism. Baton Rouge: Louisiana State University, 2006.

Forester, Jay. Industrial Dynamics. Cambridge: MIT Press, 1961. –––. Urban Dynamics. Cambridge: MIT Press, 1969. –––. World Dynamics. Cambridge: Wright-Allen Press, 1971.

Forman, Paul. “Behind Quantum Electronics: National Security as a Basis for Physical research in the United States.” Historical Studies of the Physical and Biological Sciences vol 18, no. 1 (1987): 149-229.

Fourier, Joseph. “General Remarks on the Temperatures of the Globe and the Planetary Spaces.” Translated by Ebeneser Burgess. American Journal of Science, v. 32 (1837): 1-20.

346 Freymond, Jacques. “New Dimensions in International Relations.” The Review of Politics, vol. 37, no. 4 (October, 1975): 464-478.

Friedberg, Aaron L. “Science, the Cold War, and the American State.” Diplomatic History, vol 20, no. 1 (Winter, 1996): 107-118.

Fromson, J. “A History of Federal Air Pollution Control.” Ohio State Law Journal, vol. 30, no. 3 (Summer 1969): 516-536.

Funtowicz, S.O. and J.R. Ravetz. “Three Types of Risk Assessment and the Emergence of Post-Normal Science.” In Social Theories of Risk, edited by S. Krimsky and D. Golden, 251-273. Westport, CT: Greenwood, 1993.

Gaddis, John Lewis. Strategies of Containment: A Critical Appraisal of Postwar American National Security. Oxford: Oxford University Press, 1982.

Galison, Peter and Bruce Hevly, eds. Big Science: The Growth of Large-Scale Research. Stanford, CA: Stanford University Press, 1992.

Gelbspan, Ross. Boiling Point: How Politicians, Big Oil and Coal, Journalists, and Activists Have Fueled a Climate Crisis—And What We Can Do to Avert Disaster . New York: Basic Books, 2004). –––. The Heat is On: The Climate Crisis, the Cover-up, the Prescription. New York: Basic Books, 1998.

George, Henry. Progress and Poverty: an inquiry into the cause of industrial depressions, and of increase of want with increase of wealth…the remedy. 75th Anniversary Edition. New York: Robert Schalkenbach Foundation, 1955.

Glantz, Stanton A., John Slade, Lisa A. Bero, Peter Nanauer and Deborah E. Barnes. The Cigarette Papers. Berkeley: University of California Press, 1996.

Glantz, Michael. “The Value of a Long Range Weather Forecast for the West African Sahel.” Bulletin of the American Meteorological Society, vol. 58, no.2 (February 1977): 150-158. –––. Saskatchewan Spring Wheat Production 1974: A Preliminary Assessment of a Reliable Long-range Forecast. Environment Canada Climatological Studies 33. Downsview, Ontario: Environment Canada, 1979.

Gleick, James. Chaos: Making a New Science. New York: Viking, 1987.

Gore, Albert. Earth in the Balance: Ecology and the Human Spirit. Boston: Houghton Mifflin, 1992.

347 Gottlieb, Robert. Forcing the Spring: The Transformation of the American Environmental Movement. Washington, D.C.: Island Press, 1993.

Grattan-Guiness, I. Joseph Fourier, 1768-1830: A Survey of His Life and Work. Cambridge: MIT Press, 1972.

Greenland, David. “Book Review: Climates of Hunger: Mankind and the World’s Changing Weather.” Annals of the Association of American Geographers, vol. 69, no. 2 (June, 1979): 319-320.

Guggenheim, David, director. An Inconvenient Truth. Hollywood, CA: Paramount Pictures, 2006.

Guha, Ramachandra. Environmentalism: A Global History. New York: Longman Publishing Group, 1999.

Guston, David H. “Boundary Organizations in Environmental Policy and Science: An Introduction.” Science, Technology, and Human Values, vol. 26, no. 4 (Autumn, 2001): 399-408.

Hagen, Joel. An Entangled Bank: The Origins of Ecosystem Ecology. New Brunswick: Rutgers University Press, 1992.

Hallgren, Elizabeth Lynn. The University Corporation for Atmospheric Research and the National Center for Atmospheric Research, 1960-1970: An Institutional History. Boulder, CO: National Center for Atmospheric Research, 1974.

Hammond, Allen L. “Speaking of Science: Concern Over Climate: Researchers Increasingly Go Public.” Science, vol 192, no. 4236 (April 16, 1976): 246-247.

Hansen, J., D. Johnson, A. Lacis, S. Lebedeff, P. Lee, D. Rind and G. Russell. “Climate Impact of Increasing Carbon Dioxide.” Science, vol 213, no. 4511 (August 28, 1981): 957-966.

Hare, Kenneth F. “Book Review.” Bulletin of the American Meteorological Society, vol. 58, no. 8 (August, 1976): 1016.

Harper, Kristine. Weather by the Numbers: The Genesis of Modern Meteorology. Cambridge: MIT Press, 2008.

Hart, David M. and David G. Victor. “Scientific Elites and the Making of U.S. Policy for Climate Change Research, 1957-74.” Social Studies of Science, vol. 23, no. 4 (November, 1993): 643-680.

Hartman, Dennis L. Global Physical Climatology. San Diego, CA: Academic Press, 1994.

348 Hartley, L.P. The Go Between. Middlesex: Penguin Books, 1961.

Hayden, Carl Trumbull. International Geophysical Year: Special Report Prepared by the National Academy of Sciences for the Committee on Appropriations of the United States Senate. Washington, D.C.: U.S. Government Printing Office, 1956.

Hayes, Denis. Rays of Hope: The Transition to a Post-Petroleum World. New York: W.W. Norton and Company, 1977.

Hays, Samuel P. Beauty, Health, and Permanence: Environmental Politics in the United States, 1955-1985. Cambridge University Press, 1987. –––. Conservation and the Gospel of Efficiency: The Progressive Conservation Movement, 1890-1920. Cambridge: Harvard University Press, 1959.

Head, Tom, ed. Conversations with Carl Sagan. Jackson, MS: University Press of Mississippi, 2006.

Hecht, Alan D. and Dennis Tirpak. “Framework Agreement on Climate Change: A Scientific and Policy History.” Climatic Change, vol. 29, no. 4 (April 1995): 371- 402.

Herivel, John. Joseph Fourier: The Man and the Physicist. Oxford: Clarendon Press, 1975.

Herkin, Gregg. Brotherhood of the Bomb: The Tangled Lives and Loyalties of Robert Oppenheimer, Ernest Lawrence, and Edward Teller. New York: Henry Holt, 2002. –––. “In the Service of the State: Science and the Cold War.” Diplomatic History, vol. 24, no. 1 (Winter, 2000): 107-115.

Hevly, Bruce. “Reflections on Big Science and Big History.” In Big Science: The Growth of Large-Scale Research, edited by Peter Galison and Bruce Hevly, 355-363. Stanford, CA: Stanford University Press, 1992.

Hirst, Eric. “Review: A New Prosperity: Building a Sustainable Energy Future by Solar Energy Research Institute.” Policy Sciences, vol. 14, no. 2 (April, 1982): 198-202.

Hoggan, John. Climate Cover-Up: The Crusade to Deny Global Warming. With Richard Littlemore. New York: Greystone Books, 2009.

Holden, Barry. Democracy and Global Warming. London: Continuum, 2002.

Horwitch, Mel. Clipped Wings: The American SST Conflict. Cambridge: MIT Press, 1982.

349 Hughes, Agatha and Thomas Hughes, eds. Systems, Experts, and Computers: The Systems Approach in Management and Engineering, World War II and After. Cambridge: MIT Press, 2000.

Hulme, Michael. Why We Disagree About Climate Change: Understanding Controversy, Inaction, and Opportunity. Cambridge: Cambridge University Press, 2009.

Imbrie, John and Katherine Palmer Imbrie. Ice Ages: Solving the Mystery. Cambridge: Harvard University Press, 1979.

Impact Team. The Weather Conspiracy: The Coming of the New Ice Age. New York: Ballantine Books, 1977.

International Council of Scientific Unions. Scientific Committee on Critical Problems of the Environment (SCOPE). Commission on Monitoring. Global Environmental Monitoring: A Report Submitted to the United Nations Conference on the Human Environment, Stockholm 1972. Stockholm, Sweden: ICSU-SCOPE, 1971. –––. “ICSU and Climate Science.”www.icsu.org.

Ivanova, Maria. “Looking Forward by Looking Back: Learning from UNEP’s History.” In Global Environmental Governance: Perspectives on the Current Decade, edited by Lydia Swart and Estelle Perry, 26-47. New York: Center for UN Reform Education, 2007.

Jaeger, Jill. Developing Policies for Responding to Climate Change: A Summary of the Discussions and Recommendations of the Workshops Held in Villach (28 September-2 October 1987) and Bellagio (9-13 November 1987) under the Auspices of the Beijer Insitutie, Stockholm. Geneva, Switzerland: World Meteorological Organization; Nairobi, Kenya, United Nations Environment Program, 1988.

Jasanoff, Sheila. The Fifth Branch: Science Advisors as Policymakers. Cambridge, MA: Harvard University Press, 1990. –––. “Contested Boundaries in Policy-Relevant Science,” Social Studies of Science, Vol. 17, No. 2 (May, 1987), 195-230.

Johansen, Bruce E. The Global Warming Desk Reference. Westport, Connecticut: Greenwood Press, 2002.

Johnston, Harold. “Reduction of Stratospheric Ozone by Nitrogen Oxide Catalysts from Supersonic Transport Exhaust.” Science, Vol. 173, No. 3996 (Aug., 1971), pp. 517-22. –––. “Atmospheric Ozone.” Annual Review of Physical Chemistry, vol. 43 (1992): 1-32.

Kaiser, David. American Physicists and the Cold War Bubble. Chicago: University of Chicago Press, forthcoming.

350 Kargon, Robert, Stuart W. Leslie, and Erica Schoenberger. “Far Beyond Big Science: Science Regions and the Organization of Research and Development.” In Big Science: The Growth of Large-Scale Research, edited by Peter Galison and Bruce Hevely, 334-354. Stanford, CA: Stanford University Press, 1992.

Keeling, C.D. “The Suess Effect: 13Carbon-14Carbon Interrelations.” Environment International 2 (1979): 229–300.

Kellog, W.W. and Stephen H. Schneider. “Climate Stabilization: For Better or Worse.” Science 186 (1974): 1163-1172.

Kellogg, William W. "Mankind's Impact on Climate: The Evolution of an Awareness." Climatic Change 10 (1987): 113-36.

Kennan, George F. “To Prevent a World Wasteland: A Proposal.” Foreign Affairs, vol. 48, no. 3 (April, 1970): 401-413.

Kennedy, Paul. The Parliament of Man: The Past, Present, and Future of the United Nations. New York: Random House, 2006.

Kevles, Daniel J. “Cold War and Hot Physics: Science, Security and the American State.” Historical Studies of the Physical and Biological Sciences, vol. 20, no. 2 (1990): 239-64. –––. The Physicists: The History of a Scientific Community in Modern America. Cambridge: Harvard University Press, 1995.

Kirk, Andrew. Counterculture Green: The Whole Earth Catalog and American Environmentalism. Lawrence: University Press of Kansas, 2007.

Kolbert, Elizabeth. Field Notes from a Catastrophe: Man, Nature, and Climate Change. New York: Bloomsbury, 2006.

Korsmo, Fae L. “The Genesis of the International Geophysical Year.” Physics Today (July 2007): 38-43.

Kotkin, Stephen. Armageddon Averted: The Soviet Collapse, 1970-2000. Oxford: Oxford University Press, 2008.

Kukla, George J. and R.K. Matthews. “When Will the Present Interglacial End?” Science, vol. 178, no 4057 (October 13, 1972): 190-91.

Ladurie, Emmanuel Le Roy. Times of Feast, Times of Famine: A History of Climate Since the Year 1000. Translated by Barbara Bray. New York: Noonday Press, 1971.

351 –––. “History Without People.” In The Territory of the Historian. Translated by Ben and Siân Reynolds. Chicago: University of Chicago Press, 1979.

Landsberg, Helmut E. “Climate as a Natural Resource.” The Scientific Monthly, vol. 63, no. 4 (October 1946): 293. –––. “Review: The Genesis Strategy—Climate and Global Survival.” E.O.S.: Transactions of the American Geophysical Union, vol. 57, no. 9 (September 1976): 634-5. –––. “Forum.” E.O.S.: Transactions of the American Geophysical Union, vol. 58, no. 3 (March, 1977): 122.

Lane, Earl. “Physicist Richard Garwin: A Life in Labs and the Halls of Power.” American Association for the Advancement of Science, January 17, 2006. http://www.aaas.org/news/releases/2006/0117garwin.shtml.

Lanchbery, John and David Victor. “The Role of Science in the Global Climate Negotiations.” In Green Globe Yearbook of International Co-operation and Development, edited by Helge Ole Bergesen and Georg Parmann, 29-40. Oxford: Oxford University Press, 1995.

Lash, Jonathan. A Season of Spoils: The Reagan Administration’s Attack on the Environment. New York: Pantheon Books, 1984.

Latour, Bruno. Science in Action. Cambridge: Harvard University Press, 1987. –––. The Politics of Nature: How to Bring the Sciences Into Democracy. Trans. by Catherine Porter. Cambridge: Harvard University Press, 2004.

Lauren, Paul Gordon. “Diplomats and Diplomacy of the U.N.” In The Diplomats, 1939- 1979, edited by Gordon A. Craig and Francis L. Lowenheim, 459-495. Princeton: Princeton University Press, 1994.

Leggett, Jeremy. The Carbon War: Global Warming and the End of the Oil Era. New York: Routledge, 2001.

Levine, David, ed. Defending the Earth: A Dialogue Between Murray Bookchin and Dave Foreman. Boston, MA: South End Press, 1991.

Lieber, Keir A. and Daryl Press. “The Rise of U.S. Nuclear Primacy.” Foreign Affairs, vol. 86, no. 2 (March/April, 2006): 42-55.

Lilienfeld, Robert. The Rise of Systems Theory: An Ideological Analysis. New York: John Wiley & Sons, 1978.

Lorenz, Edward. “Reflections on the Conception, Birth, and Childhood of Numerical Weather Prediction.” Annual Review of Earth and Planetary Sciences, Vol. 34 (May 2006): 37-45.

352 Lundberg, B.K.O. “Supersonic Adventure.” Bulletin of the Atomic Scientists, Vol. 21 (February, 1965), 29-33.

MacCracken, Michael C., S. B. Idso, J. Hansen, D. Johnson, A. Lacis, S. Lebedeff, P. Lee, D. Rind and G. Russell. “Climatic Effects of Atmospheric Carbon Dioxide [exchange of letters].” Science, vol. 220, no. 4599 (May 20, 1983): 873-875.

Maddox, John. The Doomsday Syndrome. New York: McGraw Hill, 1972.

Maier, Charles S. Dissolution: The Crisis of Communism and the End of East Germany. Princeton: Princeton University Press, 1997.

Malone, Thomas F. Reflections on the Human Prospect. http://humanprospect- post2.blogspot.com

Malone, Thomas F., Edward D. Goldberg, and Walter H. Monk. Roger Randall Dougan Revelle, March 7, 1909-June 15, 1991. National Academy of Sciences Biographical Memoirs. http://www.nasonline.org.

Manabe, Sukiro and Richard T. Weatherald. “Thermal Equilibrium of the Atmosphere with a Given Distribution of Relative Humidity.” Journal of the Atmospheric Sciences, vol. 24, no. 3 (May 1967): 241-259.

Manzione, Joseph. “Amusing and Amazing and Practical and Military: The Legacy of Scientific Internationalism in American Foreign Policy, 1945-1963.” Diplomatic History, vol. 24, no. 1 (Winter, 2000): 21-55.

Marsh, George Perkins. Man and Nature. 1985. 2003 edition by David Lowenthal with a new introduction by David Lowenthal and a forward by William Cronon. Seattle: University of Washington Press, 2003.

Martin, Brian. “Nuclear Winter: Science and Politics.” Science and Public Policy, vol. 15, no. 5 (October, 1988): 321-334.

Martino, Joseph Paul. Science Funding: Politics & Porkbarrel. New Brunswick, NJ: Transaction Publishers, 1992.

Mauch, Christof, Nathan Stoltzfus, and Douglas R. Weiner, eds. Shades of Green: Environmental Activism Around the Globe. Lanham, MD: Rowman & Littlefield Publishers, 2006.

McCartney, Scott. ENIAC: The Triumphs and Tragedies of the World’s First Computer. New York: Walker, 1999.

353 McCloskey, Andre T. “The Stewardship of Planet Earth: Protecting Our Planet Twenty Years of Defending the Environment,” BIOS, vol. 60, no. 1/2 (March - May, 1989): 40-43.

McCormick, John. Reclaiming Paradise: The Global Environmental Movement Bloomington: Indiana University Press, 1989.

McNamara, Robert. “Mutual Deterrence.” Speech by the Secretary of Defense, September 18, 1967, San Francisco, CA. Online at http://www.atomicarchive.com/Docs/Deterrence/Deterrence.shtml

Meadows, Donella H., Dennis L. Meadows, Jørgen Randers, and William W. Behrens, III. The Limits to Growth: A Report for the Club of Rome’s Project on the Predicament of Mankind. New York: Signet, 1972.

Meisler, Stanley. United Nations: The First Fifty Years. New York: The Atlantic Monthly Press, 1995.

Meyer, David S. A Winter of Discontent: The Nuclear Freeze Movement and American Politics. New York: Praeger, 1990.

Michaels, Patrick. Sound and Fury: The Science and Politics of Global Warming. Washinton, D.C.: Cato Institute, 1992.

Miller, Clark A. “Scientific Internationalism in American Foreign Policy: The Case of Meteorology, 1947-1958.” In Changing the Atmosphere. Expert Knowledge and Environmental Governance, edited by Clark A. Miller and Paul N. Edwards, 167- 217. Cambridge: MIT Press, 2001.

Miller, Clark A. and Paul Edwards, eds. Changing the Atmosphere: Expert Knowledge and Environmental Governance. Cambridge, MA: MIT Press, 2001.

Mitchell, J. Murray. “A Preliminary Evaluation of Atmospheric Pollution as a Cause of the Global Temperature Fluctuation of the Past Century.” In Global Effects of Environmental Pollution: Symposium on Global Environmental Pollution, edited by S. Fred Singer. New York: Springer-Verlag, 1970: 139-55.

Molina, Mario J. and F.S. Rowland. “Stratospheric Sink for Chlorofluoromethanes: Chlorine Atom-Catalyised Destruction of Ozone.” Nature 249 (June 28, 1974): 810-12.

Morgan, Judith and Neil Morgan. Roger: A Biography of Roger Revelle. San Diego: University of California, San Diego—Scripps Institution of Oceanography, 1996.

354 Naess, Arne. “Sustainable Development and Deep Ecology.” In Ethics of Environment and Development: Global Challenge, International Response. Edited by J. Ronald Engel and Joan Gibb Engel. London: John Wiley, 1992. –––. “Sustainable Development and the Deep Ecology Movement.” In Sustainable Development: Theory, Policy, and Practice Within the European Union. Edited by Susan Baker, Maria Kousis, Dick Richardson, and Steven Young, 61-71. London: Routledge, 1997.

Nash, Roderick Frazier. Wilderness and the American Mind (4th Edition). New Haven: Yale University Press, 2001.

National Academy of Sciences. Committee on Atmospheric Sciences. The Atmospheric Sciences, 1961-1971: Volume I: Goals and Plans. Washington, D.C.: National Academy of Sciences-National Research Council, 1962. –––. Committee on Atmospheric Sciences Panel on Weather and Climate Modification. Weather and Climate Modification: Problems and Prospects. Washington, D.C.: National Academy of Sciences, 1966. –––. United States Committee for the Global Atmospheric Research Program. National Research Council. Understanding Climatic Change: A Program for Action. Washington, DC: National Academy of Sciences, 1975. –––. Geophysics Research Board. Energy and Climate: Studies in Geophysics. Washington, DC: National Academy of Sciences, 1977. –––. Climate Research Board. Carbon Dioxide and Climate: A Scientific Assessment. Washington, D.C.: National Academy of Sciences, 1979. (The “Charney Report”) –––. Carbon Dioxide Assessment Committee. Changing Climate. Washington, D.C.: National Academy of Sciences, 1983.

National Academy of Sciences—National Research Council. Panel on International Meteorological Cooperation. Feasibility of a Global Observation and Analysis Experiment. Washington, D.C.: National Academy of Sciences—National Research Council, 1966.

National Defense University. Climate Change to the Year 2000: A Survey of Expert Opinion. Washington, DC, National Defense University, 1978.

National Renewable Energy Laboratory. 25 Years of Research Excellence, 1977-2002. Boulder, CO: NREL, 2002. www.nrel.gov.

National Research Council. Committee on Atmospheric Sciences. Atmospheric Ozone Studies; An Outline for an International Observation Program, A Report by the Panel on Ozone to the Committee on Atmospheric Sciences National Academy of Sciences, National Research Council, November, 1965. Washington, D.C.: National Academy of Sciences-National Research Council, 1966. –––. “Biological and Medical Effects of Nitrogen Oxide Emissions.” Chapter 3 in Environmental Impact of Stratospheric Flight: Biological and Climatic Effects of Aircraft Emissions in the Stratosphere. Report by the Climatic Impact Committee

355 for the National Academy of Sciences and the National Academy of Engineering. Washington, D.C.: National Academy of Sciences 1975. –––. Committee on Climate and Weather Fluctuations and Agricultural Production. Board on Agricultural and Renewable Resources. Climate & Food: Climatic Fluctuations and U.S. Agricultural Production. Washington, D.C.: National Academy of Sciences, 1976. –––. Committee on Impacts of Stratospheric Change. Halocarbons: Effects on Stratospheric Ozone. Washington DC: National Academy of Sciences, 1976. –––. Panel on Water and Climate. Climate, Climatic Change, and Water Supply. Washington, D.C.: National Academy of Sciences, 1977. –––. Panel on Nitrates. Nitrates: An Environmental Assessment. Washington, D.C.: National Academy of Sciences, 1978. –––. CO2/Climate Review Panel. Carbon Dioxide and Climate: A Second Assessment. Washington, DC, National Academy of Sciences, 1982.

National Science Foundation. Preliminary Plans for a National Center for Atmospheric Research: Second Progress Report of the University Committee on Atmospheric Research, February, 1959. February, 1958. Available from the National Center for Atmospheric Research. http://www.ncar.ucar.edu/publications/pubsabout.php . (also known as the “Blue Book”) –––. Federal Funds for Research and Development Detailed Historical Tables: Fiscal Years 1956-1994. NSF 94-331. Bethesda, MD: Quantum Research Corp., 1994.

Nature. “Costs and Benefits of Carbon Dioxide.” Vol. 279, no 5708 (May 3, 1979): 1.

Neimark, Peninah and Peter Rhoades Mott, eds. The Environmental Debate: A Documentary History. Westport, CT: Greenwood Press, 1999.

Nilsson, Sten and David Pitt. Protecting the Atmosphere: The Climate Change Convention and its Context. London: Earthscan Publications, 1994.

Novick, Peter. That Noble Dream: The “Objectivity Question” and the American Historical Profession. Cambridge: Cambridge University Press, 1988.

Odum, Eugene P. “The New Ecology.” BioScience, vol 14, no. 7 (July 1964), 14-16.

Oppenheimer, Michael. “Developing Policies for Responding to Climate Change.” Climatic Change, vol. 15, no. 1 (April 1989): 1-4.

Oreskes, Naomi and Ronald E. Doel. “The Physics and Chemistry of the Earth.” Chapter 28 in Cambridge History of Science, Volume V: Modern Physical and Mathematical Sciences. Edited by Mary Jo Nye, 538-560. Cambridge: Cambridge University Press, 2002.

Oreskes, Naomi and Erik M. Conway. “Challenging Knowledge: How Climate Change Became a Victim of the Cold War.” In Agnotology: The Making and Unmaking of

356 Ignorance, edited by Robert N. Proctor and Londa Schiebinger, 55-89. Stanford: Stanford University Press, 2008.

Oreskes, Naomi and Erik Conway. Merchants of Doubt: How a Handful of Scientists Obscured the Truth on Issues from Tobacco Smoke to Global Warming. New York: Bloomsbury, 2010.

Oreskes, Naomi, Erik Conway, and Matthew Shindell. “From Chicken Little to Dr. Pangloss: William Nierenberg, Global Warming, and the Social Deconstruction of Scientific Knowledge.” Historical Studies in the Natural Sciences, vol. 38, no. 1 (Winter, 2008): 109-152.

Orville, Howard T. “The Impact of Weather Control on the Cold War.” In U.S. Congress. House Committee on Interstate and Foreign Commerce. Weather Modification Research. Hearing. 85th Congress, 2nd Session, March 18-19, 1958, 51-53. Washington, D.C.: U.S. Government Printing Office, 1958.

Owen, Kenneth. Concorde and the Americans: International Politics of the Supersonic Transport. Washington: Smithsonian Institution Press, 1997.

Palmer, Trevor. Perilous Planet Earth: Catastrophes and Catastrophism through the Ages. New York: Cambridge University Press, 2003.

Panjabi, Ranee K.L. “The South and the Earth Summit: The Development/Environment Dichotomy.” In The Earth Summit at Rio: Politics, Economics, and the Environment, 93-155. Boston: Northeastern University Press, 1997.

Parker, Sybil P. McGraw-Hill Modern Scientists and Engineers. New York: McGraw Hill, 1980.

Parson, Edward A. Protecting the Ozone Layer: Science and Strategy. New York: Oxford University Press, 2003.

Paterson, Matthew. Global Warming and Global Politics. London: Routledge, 1996.

Patterson, James T. Restless Giant: The United States from Watergate to Bush v. Gore. New York: Oxford University Press, 2005.

Perkins, Sid. “Cooling Climate ‘Consensus’ of 1970s Never Was.” ScienceNews, vol. 174, no. 9 (October 25, 2008): 5-6.

Peterson, Thomas C., William M. Connolley, and John Fleck. “The Myth of the 1970s Global Cooling Scientific Consensus.” Bulletin of the American Meteorological Society, vol. 89, no. 9 (September 2008): 1325-1337.

357 Phillips, Norman. “Jule Gregory Charney, January 1, 1917-July 16, 1981.” National Academy of Sciences Biographical Memoirs. http://www.nasonline.org.

Plass, Glibert N. “Carbon Dioxide and the Climate.” American Scientist 44 (1956): 302- 316. –––. “The Carbon Dioxide Theory of Climate Change.” Tellus 8 (1956): 140-154.

Pomerance, Rafe. “The Dangers from Climate Warming: A Public Awakening.” In The Challenge of Global Warming, edited by Dean Edwin Abrahamson, 259-269. Washington, D.C.: Island Press, 1989.

Popper, Karl. Objective Knowledge: An Evolutionary Approach. Oxford: Clarendon Press, 1972.

Proctor, Robert. Value-Free Science?: Purity and Power in Modern Knowledge. Cambridge: Harvard University Press, 1991.

Pyne, Stephen. “Burning Questions and False Alarms about Wildfires at Yellowstone.” Forum for Applied Research and Public Policy, vol. 4, no. 2 (Summer 1989): 31- 39.

Rabe, Barry G. Statehouse and Greenhouse: The Emerging Politics of American Climate Change Policy. Washinton, D.C.: Brookings Institution Press, 2004.

Rasool, S.I. and S.H. Schneider. “Atmospheric Carbon Dioxide and Aerosols: Effects of Large Scale Increases on Global Climate.” Science, vol. 173, no. 3992 (July 9, 1971): 138-141.

Rasool, S. Ichtiaque. "On Predicting Calamities [with Reply]." Climatic Change, vol. 5, no. 2 (June, 1983): 201-04.

Reilly, William K. “Breakdown on the Road from Rio: Reform, Reaction and Distraction Compete in the Cause of the International Environment. 1993-1994.” Arthur and Frank Payne Lecture, Stanford University, Stanford, CA, March 2, 1994.

Revelle, Roger. Marine Bottom Samples Collected in the Pacific Ocean by the Carnegie on its Seventh Cruise. Ph.D. dissertation, University of California, 1936. –––. “Introduction: The Scientific History of Carbon Dioxide.” In The Carbon Cycle and Atmospheric CO2: Natural Variations Archean to Present, edited by E.T. Sundquist and W.S. Broecker. Geophysical Monograph 32. Washington, D.C.: American Geophysical Union, 1985.

Revelle, Roger and Hans E. Suess. “Carbon Dioxide Exchange between Atmosphere and Ocean and the Question of an Increase of Atmospheric CO2 During the Past Decades.” Tellus 9 (1957): 18-27.

358 Rhodes, Richard. The Making of the Atomic Bomb. New York: Simon and Schuster, 1986.

Roan, Sharon L. Ozone Crisis: The 15 Year Evolution of a Sudden Global Emergency. New York: Wiley and Sons, 1989.

Robertson, Thomas. “‘This is the American Earth’: American Empire, the Cold War, and American Environmentalism.” Diplomatic History, vol. 32, no. 24 (September, 2008): 561-584.

Rochon, Thomas R. and David S. Mayer. Coalitions and Political Movements: The Lessons of the Nuclear Freeze. Boulder, CO: Lynne Rienner, 1997.

Rosenbaum, E.M. with W. Hoffer. Betrayal: The Untold Story of the Kurt Waldheim Investigation and Cover-Up. New York: St. Martin's Press, 1993.

Rosenbloom, Joshua. “The Politics of the American SST Programme: Origin, Opposition, and Termination.” Social Studies of Science, Vol. 11, No. 4 (Nov. 1981), 403-423.

Rothman, Hal K. The Greening of a Nation?: Environmentalism in the United States Since 1945. Fort Worth, TX: Harcourt Brace & Company, 1998.

Rozwadowski, Helen. The Sea Knows No Boundaries: A Century of Marine Science Under ICES. Seattle: University of Washington Press, 2002. –––. Fathoming the Ocean: The Discovery and Exploration of the Deep Sea. Cambridge: Harvard University Press, 2005.

Rusin, N. and L. Flit. Man Versus Climate. Translated by Dorian Rottenberg. Moscow: Peace Publishers, 1960.

Sagan, Carl. “Nuclear War and Climate Catastrophe: Some Policy Implications.” Foreign Affairs, vol. 62, no. 2 (Winter, 1983-84): 257-92.

Sale, Kirkpatrick. The Green Revolution: The American Environmental Movement, 1962- 1992. New York: Hill & Wang, 1993.

Savage, James. Funding Science in America: Congress, Universities, and the Politics of the Academic Pork Barrel. New York: Cambridge University Press, 1999.

Schaller, Michael. Reckoning with Reagan: America and Its President in the 1980s. Oxford: Oxford University Press, 1980.

Schell, Jonathan. The Fate of the Earth. New York: Random House, 1982.

Schneider, Stephen H. “On the Carbon Dioxide-Climate Confusion.” Journal of the Atmospheric Sciences, vol. 32, issue 11 (November, 1975): 2060-2066.

359 –––. The Genesis Strategy: Climate and Global Survival. With Lynn E. Mesirow. New York: Plenum Press, 1976. –––. The Primordial Bond : Exploring Connections Between Man and Nature Through the Humanities and Sciences. New York: Plenum Press, 1981. –––. Global Warming: Are We Entering a Greenhouse Century? San Franscisco: Sierra Club Books, 1989. –––. Laboratory Earth: The Planetary Gamble We Can’t Afford to Lose. New York: Basic Books, 1997. –––. Science as a Contact Sport: Inside the Battle to Save Earth’s Climate. Washington, D.C.: National Geographic, 2009.

Schneider, Stephen. http://stephenschneider.stanford.edu.

Schneider, Stephen H., William W. Kellogg, V. Ramanathan, Conway B. Leovy and Sherwood B. Idso. “Carbon Dioxide and Climate [exchange of letters].” Science, vol. 210, no. 4465 (October 3, 1980): 6-8.

Schneider, Stephen H. and Randi Londer. The Coevolution of Climate and Life. San Francisco: Sierra Club Books, 1984.

Schneider, Stephen H., Armin Rosencranz, and John O. Niles, eds. Climate Change Policy: A Survey. Washington, D.C.: Island Press, 2002.

Scott, James C. Seeing Like a State: How Certain Schemes to Improve the Human Condition Have Failed. New Haven: Yale University Press, 1998.

Sears, Paul B. (1937) Deserts on the March. 4th Ed. Norman: University of Oklahoma Press, 1980.

Seitz, Russel. Letter. Foreign Affairs, vol. 63, no. 4 (Spring, 1984): 998-999.

Seidel, Stephen and Dale Keyes. Can We Delay a Greenhouse Warming? 2nd ed. Washington, D.C.: Environmental Protection Agency, 1983

Shabecoff, Philip. A Fierce Green Fire: The American Environmental Movement (revised edition). Washington, D.C.: Island Press, 2003.

Shackley, Simon and Brian Wynne. “Representing Uncertainty in Global Climate Change Science and Policy.” Science, Technology, and Human Values, vol. 21, no. 3 (Summer, 1996): 275-302

Shiva, Vandana. The Violence of Green Revolution: Third World Agriculture, Ecology, and Politics. London: Atlantic Highlands, 1991. –––. “The Greening of the Global Reach." In Global Ecology: A New Arena for Political Conflict. Edited by Wolfgang Sachs, 249-56. London: Zed Books, 1993.

360 Singer, S.F. “Is the ‘Nuclear Winter’ Real?” Nature 310 (August 16, 1984): 625.

Singer, S. Fred, Cresson H. Kearny, R. P. Turco, O. B. Toon, T. P. Ackerman, J. B. Pollack and C. Sagan. “On a ‘Nuclear Winter.’” Science, vol. 227, no. 4685 (Jan. 25, 1985): 356-444.

Smagorinski, Joseph. “Global Atmospheric Modeling and the Numerical Simulation of Climate.” In Weather and Climate Modification, edited by Wilmot Hess. New York: Wiley, 1974.

Smith, Jeffrey. “‘Nuclear Winter’ Feels Budgetary Chill.” Science, vol. 227, 4689 (February 22, 1985): 294-5.

Solar Energy Research Institute. A New Prosperity: Building a Sustainable Future. Andover, MA: Brickhouse Publishing, 1981.

Spangenburg, Ray and Kit Moser. Carl Sagan: A Biography. Westport, CT: Greenwood Press, 2004.

Speth, James Gustave (ed). Worlds Apart: Globalization and the Environment. Washington, D.C: Island Press, 2003. –––. Red Sky at Morning: American and the Crisis of the Global Environment. New Haven: Yale University Press, 2004.

Speth, James Gustave and Peter M. Haas. Global Environmental Governance. Washington, D.C.: Island Press, 2006.

Stanhill, Gerald. “Climate Change Science is now Big Science.” Eos, Transactions of the American Geophysical Union 80, no. 35 (1999): 396.

Swedin, Eric G. and David L. Ferro. Computers: The Life Story of a Technology. Baltimore: Johns Hopkins University Press, 2007.

Stevens, William K. The Change in the Weather: People, Weather, and the Science of Climate. New York: Dell Publishing, 1999.

Study of Man’s Impact on Climate. Inadvertent Climate Modification: Report of the Study of Man’s Impact on Climate. Cambridge: MIT Press, 1971.

Study of Critical Environmental Problems. Man’s Impact on the Global Environment. Assessment and Recommendation for Action: Report on the Study of Critical Environmental Problems. Cambridge: MIT Press, 1970.

Suess, Hans E. “Radiocarbon Concentration in Modern Wood.” Science, vol. 122, no 3166 (2 September 1955): 415-17.

361 Sullivan, Walter. Assault on the Unknown: The International Geophysical Year. New York: McGraw Hill, 1961.

Swedish Preparatory Committee for the U.N. Conference on the Human Environment (Bert Bolin, Chairman). Sweden’s Case Study for the United Nations Conference on the Human Environment: Air Pollution Across National Boundaries. Stockholm: Norstadt and Sons, 1972.

Teller, Edward. “Widespread After-Effects of Nuclear War.” Nature 310 (August 16, 1984): 621-62.

Thomas, William Leroy, ed., Man’s Role in Changing the Face of the Earth. Chicago: University of Chicago Press, 1956.

Thompson, Louis M. “Weather Variability, Climatic Change, and Grain Production.” Science, vol. 188, no. 4188 (May 9, 1975): 535-541.

Thompson, Starley L. and Stephen H. Schneider. “Nuclear Winter Reappraised.” Foreign Affairs, vol. 64, no. 5 (Summer, 1986): 981-1005.

Trager, James. Amber Waves of Grain: The Secret Russian Wheat Sales That Sent American Food Prices Soaring. New York: Arthur Fields Books, 1973. –––. The Great Grain Robbery. New York: Ballantine Books, 1975.

Turco, Richard P. “Carl Sagan and Nuclear Winter.” In Carl Sagan’s Universe, edited by Yervant Terzian and Elizabeth Bilson, 239-46. New York: Cambridge University Press, 1997.

Turco, R., O. B. Toon, T. Ackerman, J. B. Pollack, and C. Sagan. “Nuclear Winter: Global Consequences of Multiple Nuclear Explosions.” Science, vol. 222, no. 4630 (December 23, 1983): 1283-92.

Turner, Fred. From Counterculture to Cyberculture: Stewart Brand, the Whole Earth Network, and the Rise of Digital Utopianism. Chicago: University of Chicago Press, 2008.

Tyndall, John. “On the Absorption and Radiation of Heat by Gases and Vapours, and on the Physical Connection of Radiation, Absorption, and Conduction.” Philosophical Magazine, ser. 4, 22 (1861): 169-94, 273-85 –––. “On Radiation Through the Earth’s Atmosphere.” Philosophical Magazine, ser. 4, 125 (1862): 200-206.

United Nations Environment Program. “Declaration of the United Nations Conference on the Human Environment.” Programme.” June 16, 1972. http://www.unep.org/Documents.Multilingual/Default.asp?DocumentID=97&Arti cleID=1503.

362 Ungar, Sheldon. “The Rise and (Relative) Decline of Global Warming as a Social Problem.” Sociological Quarterly, vol. 33, no. 4 (December, 1992): 483-501.

U.S. Congress. House of Representatives. Committee on Appropriations. Hearing. Report on the International Geophysical Year. 85th Congress, 1st Session, May 1, 1957. Washington, D.C.: Government Printing Office, 1957. –––. House Committee on Interstate and Foreign Commerce. Hearing. Weather Modification Research. 85th Congress, 2nd Session, March 18-19, 1958. Washington, D.C.: U.S. Government Printing Office, 1958. –––. House of Representatives. Committee on Appropriations. Hearing. Civil Supersonic Aircraft Development (SST). 92nd Congress, 1st Session, March 1-4, 1971. Washington, D.C.: U.S. Government Printing Office, 1971. –––. House of Representatives. Committee on Government Operations. Subcommittee on Government Activities and Transportation Hearing. Hearing. FAA Certification of the SST Concorde. 94th Congress, 1st Session, July 24, Nov. 13-14, Dec. 9 and 12, 1975, and Feb. 24, 1976. Washington, D.C.: U.S. Government Printing Office, 1975. –––. House of Representatives. Committee on Science and Technology. Hearing. The National Climate Program Act. 94th Congress, Second Session, May 18-20, 25- 27, 1976. Washington, D.C.: U.S. Government Printing Office, 1976. –––. House of Representatives. Committee on Science and Technology. Hearing. Effects of Carbon Dioxide Buildup in the Atmosphere. 96th Congress, 2nd Session, April 3, 1980. Washington, D.C.: U.S. Government Printing Office, 1980. –––. House of Representatives. Committee on Interior and Insular Affairs. Briefing by the Secretary of the Interior. 97th Congress, 1st Session, February 5, 1981. Washington, D.C.: U.S. Government Printing Office, 1981. –––. House of Representatives. Committee on Appropriations. Hearing. Department of the Interior and Related Agencies Appropriations for 1982, Part 5. 97th Congress, 1st Session, February 25, March 4, 6, 10-12, 16, 20, and April 7, 1981. Washington, D.C.: U.S. Government Printing Office, 1981. –––. House of Representatives. Committee on Science and Technology. Hearing. Carbon Dioxide and Climate: The Greenhouse Efffect. 97th Congress, 1st Session, July 31, 1981. Washington, D.C.: U.S. Government Printing Office, 1981. –––. House of Representatives. Committee on Science and Technology. Hearing. Environmental Protection Agency Research and Development Posture. 97th Congress, 1st Session, October 22, 1981. Washington, D.C.: U.S. Government Printing Office, 1981. –––. House of Representatives. Committee on Appropriations. Hearing. Secretary Watt’s Refusal to Cooperate with Congress. 97th Congress, 2nd Session, February 24, 1982. Washington, D.C.: U.S. Government Printing Office, 1982. –––. House of Representatives. Committee on Science and Technology. Hearing. Oversight of the National Climate Program. 97th Congress, 1st Session, March 5, 1981 and March 16, 1982. Washington, D.C.: U.S. Government Printing Office, 1982.

363 –––. House of Representatives. Committee on Science and Technology. Hearing. Fiscal Year 1983 Department of Energy Budget Review (Fossil Energy and Basic Research). 97th Congress, 2nd Session, March 18, 23, 24, 25, 30, 1982. Washington, D.C.: U.S. Government Printing Office, 1982. –––. House Committee on Science and Technology. Hearing. Carbon Dioxide and Climate: The Greenhouse Effect. 97th Congress, 2nd Session, March 25, 1982. Washington, D.C.: U.S. Government Printing Office, 1982. –––. House of Representatives. Committee on Science and Technology. Carbon Dioxide and the Greenhouse Effect. 98th Congress, 2nd Session, February 28, 1984. Washington, D.C.: U.S. Government Printing Office, 1984. –––. House of Representatives. Committee on Interior and Insular Affairs. Implications of Global Warming for Natural Resources. Hearing. 100th Congress, 2nd Session, September 27th in Washington, D.C. and October 17th in San Francisco, CA, 1988. Washington, D.C.: U.S. Government Printing Office, 1989. –––. Joint Economic Committee. Subcommittee on Economy in Government. Hearing. Economic Analysis and the Efficiency of Government. Part 4: Supersonic Transport Development. 91st Congress, 2nd Session, May 7, 11, and 12, 1970. Washington, D.C.: U.S. Government Printing Office,1970. –––. Joint Economic Committee. Hearing. The Supersonic Transport. 92nd Congress, 2nd Session, December 27-28, 1972. Washington, D.C.: U.S. Government Printing Office, 1972. –––. Senate. Committee on Foreign Relations. U.N Conference on the Human Environment: Preparations and Prospects. 92nd Congress, 2nd Session. May 3-5, 1972. Washington, D.C.: U.S. Government Printing Office, 1972. –––. Senate. Committee on Foreign Relations. Hearing. Weather Modification. 93rd Congress, 2nd Session. January 25 and March 20, 1974. Washington, D.C.: U.S. Government Printing Office, 1974. –––. Senate. Committee on Commerce, Science, and Transportation. Hearings. National Climate Program Act. 95th Congress, 1st Session, June 8-10 and July 5, 6, and 8, 1977. Washington, D.C.: U.S. Government Printing Office, 1977. –––. Senate. Committee on Energy and Natural Resources. James G. Watt Nomination. 97th Congress, 1st Session, January 7 and 8, 1981. Washington, D.C.: U.S. Government Printing Office, 1981. –––. Senate. Committee on Environment and Public Works. Global Warming. Hearing. 99th Congress, 1st Session, December 10, 1985. Washington, D.C.: U.S. Government Printing Office, 1986. –––. Senate. Committee on Environment and Public Works. Ozone Depletion, the Greenhouse Effect, and Climate Change. Hearings, 99th Congress, 2nd Session, June 10 and 11, 1986. Washington, D.C.: U.S. Government Printing Office, 1986. –––. Senate. Committee on Agriculture, Nutrition, and Forestry. Hearing. Global Climate Change Prevention Act of 1989. 101st Congress, 1st Session, November 6, 1989. Washington, D.C., U.S. Government Printing Office, 1989.

U.S. Environmental Protection Agency. “Government Ban on Fluorocarbon Gases in Aerosol Products Begins October 15 [1978].” EPA Press Release, October 15, 1978. http://www.epa.gov/history/topics/ozone/01.htm.

364 U.S. President’s Science Advisory Committee. Environmental Pollution Panel. Restoring the Quality of Our Environment: Report of the Pollution Panel, President’s Science Advisory Committee. Washington, D.C.: The White House, 1965.

Van der Sluijs, Jeroen, Josee van Eijndhoven, Simon Shackley, and Brian Wynne. “Anchoring Devices in Science for Policy: The Case of Consensus around Climate Sensitivity.” Social Studies of Science, vol. 28, no. 2 (April, 1998): 291- 323.

Visser, Dana R. and Guillermo A. Mendoza. “Debt-for-Nature Swaps in Latin America.” Journal of Forestry, vol. 92, no. 6 (November, 1994): 13-16.

Wade, Nicholas. “CO2 in Climate: Gloomsday Predictions Have No Fault.” Science, vol. 206, no. 4421 (November 23, 1979): 912-913.

Waenke, Heinrich and James R. Arnold. “Hans E. Suess, 1909-1993.” National Academy of Sciences Biographical Memoirs. http://www.nasonline.org/site/PageServer?pagename=MEMOIRS_S.

Walker, Tim E. “Promoting Air Quality, Resisting Climate Change: The George H.W. Bush Administration and the Economic Precautionary Principle.” Paper presented to the Annual Conference of the American Society of Environmental Historians, Boise, Idaho, March 15, 2008.

Wallace, Linda L., ed. After the Fires: The Ecology of Change in Yellowstone National Park. New Haven: Yale University Press, 2004.

Waller, Douglas C. Congress and the Nuclear Freeze: An Inside Look at the Politics of a Mass Movement. Amherst, MA: University of Massachusetts Press, 1987.

Wang, Zuoyue. In Sputnik’s Shadow: The President’s Science Advisory Committee and Cold War America. New Brunswick, NJ: Rutgers University Press, 2008.

Ward, Barbara. Spaceship Earth. New York: Columbia University Press, 1968.

Washington, Warren M. Odyssey in Climate Modeling, Global Warming, and Advising Five Presidents. Morrisville, NC: Lulu.com, 2006.

Westad, Odd Arne. The Global Cold War: Third World Interventions and the Making of Our Time. New York: Cambridge University Press, 2005.

Williams, Jerry L. The Rise and Decline of Public Interest in Global Warming: Toward a Pragmatic Conception of Environmental Problems. Huntington, New York: Nova Science Publishers, Inc., 2001.

365 Weart, Spencer. The Discovery of Global Warming. Cambridge: Harvard University Press, 2003. –––. The Discovery of Global Warming: A Hypertext History of how Scientists Came to (Partly)Understand What People are Doing to Cause Climate Change. www.aip.org/history/climate/index.html. –––. “Climate Modification Schemes.” The Discovery of Global Warming: A Hypertext History of how Scientists Came to (Partly)Understand What People are Doing to Cause Climate Change. http://www.aip.org/history/climate/RainMake.htm.

Weinberg, Alvin. “Impact of Large-Scale Science on the United States.” Science, vol. 134, no. 3473 (July, 1961): 161-64. –––. Reflections on Big Science. Cambridge: MIT Press, 1967.

Weinberger, Caspar. Fighting for Peace: Seven Critical Years in the Pentagon. New York: Warner Books, 1990.

Weinberger, Caspar W. “The Potential Effects of Nuclear War on the Climate: A Report to the United States Congress.” March, 1985. In U.S. Congress. Senate. Armed Services Committee. Hearing. Nuclear Winter and its Implications. 99th Congress, First Session, October 2 and 3, 1985, 72-90. Washington, D.C.: U.S. Government Printing Office, 1985.

Weiner, Douglas. A Little Corner of Freedom: Russian Nature Protection from Stalin to Gorbachev. Berkeley: University of California Press, 1999.

Weiner, Jonathan. The Next One Hundred Years: Shaping the Future of Our Living Earth. New York: Bantam, 1990.

West, Heather. “Agriculture, Climate Change, and Ronald Reagan.” Working paper, Stanford University, June 8, 2007.

Whitaker, John C. Striking a Balance: Environment and Natural Resources Policy in the Nixon-Ford Years. Washington, D.C.: American Enterprise Institution for Public Policy Research, 1976.

Wilson, Andrew. The Concorde Fiasco. Middlesex: Penguin Books, 1973.

Wilson, Carroll L. and William H. Matthews, eds. Inadvertent Climate Modification. Report of Conference, Study of Man’s Impact on Climate (SMIC), Stockholm. Boston: MIT Press, 1971.

Wohl, C.G. “Scientist as Detective: Luis Alvarez and the Pyramid Burial Chambers, the JFK Assassination, and the End of the Dinosaurs.” American Journal of Physics, vol. 75, no. 11 (November, 2007): 968-977.

366 Woodwell, George. “The Carbon Dioxide Question.” Scientific American, vol. 238, no. 1 (January, 1978): 34-43.

Woodwell, G.M., R.H. Whittaker, W.A. Reiners, G.E. Likens, C.C. Delwiche, and D.B. Botkin. “The Biota and the World Carbon Budget.” Science, vol. 199, no 4325 (January 13, 1978): 141-146.

World Climate Program. Report of the International Conference on the Assessment of the Role of Carbon Dioxide and of Other Greenhouse Gases on Climate Variations and Associated Impacts. WMO No. 661. Geneva, Switzerland: World Meteorological Organization, 1986.

World Commission on Environment and Development. Our Common Future: Report on the World Commission on Environment and Development. Oxford: Oxford University Press, 1987.

World Meteorological Organization. Proceedings of the World Climate Conference—A Conference of Experts on Climate and Mankind, 12-23 February, 1979. WMO No. 537. Geneva: World Meteorological Organization, 1979. –––. Report of the International Conference on the Assessment of the Role of Carbon Dioxide and of Other Greenhouse Gases in Climate Variations and Associated Impacts, Villach, Austria, 9-15 October, 1985. WMO No. 661. Geneva: World Meteorological Organization, 1986. –––. Proceedings of the World Conference on the Changing Atmosphere: Implications for Global Security, Toronto, Canada, June 27–30. WMO No. 710. Geneva: WMO/UNEP, 1988.

Yamarella, Ernest Y. and Randal H. Ihara, eds. The Acid Rain Debate: Scientific, Economic, and Political Dimensions. Boulder, CO: Westview Press, 1985.

367