Effects of Human-Caused Greenhouse Gas Emissions On U.S. Climate, Health, and Welfare (2025)
Effects of Human-Caused Greenhouse Gas Emissions On U.S. Climate, Health, and Welfare (2025)
org/29239
DETAILS
136 pages | 8.5 x 11 | PAPERBACK
ISBN 978-0-309-99603-7 | DOI 10.17226/29239
CONTRIBUTORS
Committee on Anthropogenic Greenhouse Gases and U.S. Climate: Evidence and
Impacts; Climate Crossroads; Board on Atmospheric Sciences and Climate; Board
BUY THIS BOOK on Health Sciences Policy; National Research Council Executive Office; Division
on Earth and Life Studies; Health and Medicine Division; National Academies of
Sciences, Engineering, and Medicine
Visit the National Academies Press at nap.edu and login or register to get:
– Access to free PDF downloads of thousands of publications
– 10% off the price of print publications
– Email or social media notifications of new titles related to your interests
– Special offers and discounts
All downloadable National Academies titles are free to be used for personal and/or non-commercial
academic use. Users may also freely post links to our titles on this website; non-commercial academic
users are encouraged to link to the version on this website rather than distribute a downloaded PDF
to ensure that all users are accessing the latest authoritative version of the work. All other uses require
written permission. (Request Permission)
This PDF is protected by copyright and owned by the National Academy of Sciences; unless otherwise
indicated, the National Academy of Sciences retains copyright to all materials in this PDF with all rights
reserved.
Effects of Human-Caused Greenhouse Gas Emissions on U.S. Climate, Health, and Welfare
Climate Crossroads
This study was supported by the National Academy of Sciences Arthur L. Day Fund and Ralph J.
Cicerone and Carol M. Cicerone Endowment for NAS Missions. Any opinions, findings,
conclusions, or recommendations expressed in this publication do not necessarily reflect the
views of any individual or organization that provided support for the project.
This publication is available from the National Academies Press, 500 Fifth Street, NW, Keck
360, Washington, DC 20001; (800) 624-6242; https://nap.nationalacademies.org.
The manufacturer’s authorized representative in the European Union for product safety is
Authorised Rep Compliance Ltd., Ground Floor, 71 Lower Baggot Street, Dublin D02 P593
Ireland; www.arccompliance.com.
Suggested citation: National Academies of Sciences, Engineering, and Medicine. 2025. Effects of
Human-Caused Greenhouse Gas Emissions on U.S. Climate, Health, and Welfare. Washington,
DC: National Academies Press. https://doi.org/10.17226/29239.
Prepublication Copy
The National Academy of Sciences was established in 1863 by an Act of Congress, signed by
President Lincoln, as a private, nongovernmental institution to advise the nation on issues
related to science and technology. Members are elected by their peers for outstanding
contributions to research. Dr. Marcia McNutt is president.
The National Academy of Engineering was established in 1964 under the charter of the
National Academy of Sciences to bring the practices of engineering to advising the nation.
Members are elected by their peers for extraordinary contributions to engineering. Dr. Tsu-Jae
Liu is president.
The National Academy of Medicine (formerly the Institute of Medicine) was established in
1970 under the charter of the National Academy of Sciences to advise the nation on medical and
health issues. Members are elected by their peers for distinguished contributions to medicine
and health. Dr. Victor J. Dzau is president.
The three Academies work together as the National Academies of Sciences, Engineering,
and Medicine to provide independent, objective analysis and advice to the nation and conduct
other activities to solve complex problems and inform public policy decisions. The National
Academies also encourage education and research, recognize outstanding contributions to
knowledge, and increase public understanding in matters of science, engineering, and medicine.
Learn more about the National Academies of Sciences, Engineering, and Medicine at
www.nationalacademies.org.
Prepublication Copy
For information about other products and activities of the National Academies, please
visit www.nationalacademies.org/about/whatwedo.
Prepublication Copy
Study Staff
Prepublication Copy v
Reviewers
This Consensus Study Report was reviewed in draft form by individuals chosen for their diverse
perspectives and technical expertise. The purpose of this independent review is to provide candid and
critical comments that will assist the National Academies of Sciences, Engineering, and Medicine in
making each published report as sound as possible and to ensure that it meets the institutional standards
for quality, objectivity, evidence, and responsiveness to the study charge. The review comments and draft
manuscript remain confidential to protect the integrity of the deliberative process.
We thank the following individuals for their review of this report:
Although the reviewers listed above provided many constructive comments and suggestions, they
were not asked to endorse the conclusions or recommendations of this report nor did they see the final
draft before its release. The review of this report was overseen by KAI N. LEE, Center for Ocean
Solutions, Stanford University, and Owl of Minerva LLC, and CYNTHIA M. BEALL (NAS), Case
Western Reserve University. They were responsible for making certain that an independent examination
of this report was carried out in accordance with the standards of the National Academies and that all
review comments were carefully considered. Responsibility for the final content rests entirely with the
authoring committee and the National Academies.
Contents
FOREWORD .................................................................................................................................................... xi
PREFACE........................................................................................................................................................ xiii
SUMMARY ........................................................................................................................................................ 1
1 INTRODUCTION ............................................................................................................................................ 3
1.1 Study Charge, 3
1.2 Report Structure, 4
1.3 Committee’s Approach, 6
1.4 Geographic Focus of the Report, 7
1.5 Human Health and Public Welfare Impacts Considered, 7
Prepublication Copy ix
x Contents
REFERENCES ................................................................................................................................................ 72
APPENDIXES
Prepublication Copy
Foreword
The U.S. Environmental Protection Agency (EPA) concluded that “six greenhouse gases taken in
combination endanger both the public health and the public welfare of current and future generations” in
its 2009 “Endangerment Finding.” In a Federal Register Notice published on August 1, 2025, EPA stated
that the agency “unreasonably analyzed the scientific record” in making the 2009 Endangerment Finding
and that subsequent “developments cast significant doubt on the reliability of the findings.” Such
significant claims about the scientific record deserve careful review. The Federal Register Notice
proposed repealing the Endangerment Finding and invited public comments.
In response to EPA’s request for public input, the National Academies of Sciences, Engineering,
and Medicine undertook this independent assessment of the science underpinning the Endangerment
Finding. In the Clean Air Act, the U.S. Congress instructed EPA to draw on findings, recommendations,
and comments from the Clean Air Science Advisory Committee (CASAC) and the National Academy of
Sciences (NAS) (42 U.S. Code § 7607(D)(3)(c)). Advice from CASAC was not available to EPA during
the window when it was considering this proposed rulemaking because CASAC was disbanded in
January 2025 and EPA was in the process of appointing new members (FR Doc. 2025-07538 (90 FR
18658)).
Supporting evidence-informed decision-making by the federal government is a core mission of
the National Academies. This study was produced to meet the timeline of the Federal Register Notice
(August 1-September 22, 2025) and followed the standard National Academies’ processes for managing
conflicts of interest, inviting public comment on the committee members, and thorough peer review of the
draft report.
This study was supported by the NAS, using funding from two of its endowments. The Arthur L.
Day Fund was created for studies on physics of the Earth; its namesake was an expert in geophysics and
volcanology who served as vice president of the NAS from 1933-1941. The Ralph J. Cicerone and Carol
M. Cicerone Fund was created to honor the service of Dr. Ralph Cicerone, president of the NAS from
2005-2016.
I am deeply grateful to Dr. Shirley Tilghman, who ably chaired this committee, and to all of the
members and staff who worked tirelessly to complete this report in a timely manner.
Marcia McNutt
President
National Academy of Sciences
Prepublication Copy xi
Preface
As the committee undertook this project, it was hard not to think about recent climate-related
disasters: the heavy rainfall of Hurricane Helene that destroyed homes and roads in the mountains of
North Carolina, the fast-moving wildfires that displaced thousands in Los Angeles and affected air quality
for miles around, and the rapid flooding of the Guadalupe River in central Texas that led to at least 135
fatalities. The U.S. Environmental Protection Agency (EPA) concluded in 2009, based on scientific
understanding at the time, that emitting greenhouse gases (GHGs) to the atmosphere increased the risk of
harms to human health and welfare from changes to the climate, including the risks associated with
hurricanes, wildfires, and heavy rainfall, among many others. On the strength of this “Endangerment
Finding,” EPA and many state governments instituted their own regulations governing GHG emissions in
the intervening years.
In August 2025 EPA issued a notice of proposed rulemaking to rescind the Endangerment
Finding. This 2009 finding by the EPA Administrator was informed by a companion Technical Support
Document (EPA, 2009) that laid out the scientific evidence that emissions of six GHGs posed a threat to
human health and welfare. With the aim of informing EPA as it considers the status of the Endangerment
Finding, the National Academies undertook this study to evaluate the current state of scientific evidence
regarding the impact of human-caused GHGs on climate, with a particular focus on the evidence in the
peer-reviewed primary literature that has accumulated since 2009.
Specifically, the committee asked whether new evidence since 2009 strengthened or weakened
the primary conclusions in EPA (2009) and addressed uncertainties that remain in our understanding of
the science of climate change. In addition, the committee identified new issues that were not evident or
addressed in EPA (2009). Although climate change in response to GHG emissions is a global issue, the
committee concentrated on the effects on human health and welfare in the United States in order to
address the statutory concern of EPA. The committee’s charge was to produce a succinct and balanced
evaluation of the state of the science, not to make recommendations or advocate for a specific policy. We
hope that the report will serve as a critical resource in informing U.S. federal agency decision-making
regarding future GHG regulations.
The importance of getting the science right weighed heavily on the committee’s deliberations,
given the potential significant implications of a changing climate and of the actions proposed to address
it. Unlike earthquakes and volcanoes, over which we have no control, responding to the potential harm to
human health and welfare from changes in the climate is actionable now. While the short timeline of the
study did not lend itself to holding open discussion sessions in person, the committee is grateful to the
more than 200 individuals and organizations who responded to a Request for Information. These inputs
helped the committee survey the breadth of the literature that has been published since 2009, pointing it to
more than 600 peer-reviewed articles. Many of those contributions have been incorporated into the report
and influenced its conclusions.
I am deeply grateful to the 15 distinguished scientists, engineers, and physicians on the committee,
who so generously gave their time and expertise to produce the report. They would not have succeeded
without the logistical, managerial, and editorial support of the National Academies’ staff—Katherine
Bowman, Nancy Huddleston, April Melvin, Lindsay Moller, Maddi Nicol, Amanda Purcell, and Kasey
White, ably led by Amanda Staudt. Everyone, including those who participated in the report review process,
was inspired by the importance of the task at hand, and contributed significantly to the final report.
Shirley M. Tilghman, Chair
Committee on Anthropogenic Greenhouse Gases
and U.S. Climate: Evidence and Impacts
September 2025
Summary1
The scientific community has been studying the question of how human-caused emissions of
greenhouse gases are affecting the climate for well over a century. Much is known today, drawing upon
decades of direct observations of the Earth system and detailed research. In this report, the committee
summarizes the latest evidence on whether greenhouse gas (GHG) emissions threaten human health and
welfare in the United States.
The impetus for this report was a notice of proposed rulemaking issued by the U.S.
Environmental Protection Agency (EPA) indicating its intention to rescind the 2009 Endangerment and
Cause or Contribute Findings for Greenhouse Gases Under Section 202(a) of the Clean Air Act.
Recognizing that significantly more evidence is available today, the National Academies of Sciences,
Engineering, and Medicine launched this study to review newly available scientific evidence on the topics
considered in the Technical Support Document that EPA prepared in considering whether to make the
finding (see Box 1.2 for the Statement of Task). The committee’s report focuses on evidence gathered by
the scientific community since the Technical Support Document was published in 2009 and describes
supporting evidence, the level of confidence, and areas that are under continuing debate or are unknown.
On the basis of the scientific evidence outlined in the body of this report, the committee reached
the following overarching conclusion:
(1) Emissions of greenhouse gases from human activities are increasing the concentration of
these gases in the atmosphere. Human activities, such as the extraction and burning of fossil
fuels, cement and chemical production, deforestation, and agricultural activities, emit GHGs,
which include carbon dioxide (CO2), methane (CH4), nitrous oxide (N2O), and fluorinated gases
(F-gases), to the atmosphere. Total global GHG emissions continue to increase, even though U.S.
emissions of CO2 have decreased slightly in recent years largely due to changes in energy
production and consumption. Multiple lines of evidence show that GHG emissions from human
activities are the primary driver of the observed long-term warming trend. No known natural
drivers, such as incoming solar radiation or volcanic emissions, can explain observed changes.
(2) Improved observations confirm unequivocally that greenhouse gas emissions are
warming Earth’s surface and changing Earth’s climate. Longer records, improved and more
robust observational networks, and analytical and methodological advances have strengthened
detection of observed changes and their attribution to elevated GHGs. Trends observed include
increases in hot extremes and extreme single-day precipitation events, declines in cold extremes,
regional shifts in annual precipitation, warming of the Earth’s oceans, a decrease in ocean pH,
rising sea levels, and an increase in wildfire severity.
1
This summary does not include references. Citations for the information presented herein are provided in the main
text.
Prepublication Copy 1
(3) Human-caused emissions of greenhouse gases and resulting climate change harm the
health of people in the United States. Climate change intensifies risks to humans from
exposures to extreme heat, ground-level ozone, airborne particulate matter, extreme weather
events, and airborne allergens, affecting incidence of cardiovascular, respiratory, and other
diseases. Climate change has increased exposure to pollutants from wildfire smoke and dust,
which has been linked to adverse health effects. The increasing severity of some extreme events
has contributed to injury, illness, and death in affected communities. Health impacts related to
climate-sensitive infectious diseases—such as those carried by insects and in contaminated
water—have increased. New evidence is developing about additional health impacts of climate
change, including on mental health, nutrition, immune health, antimicrobial resistance, kidney
disease, and negative pregnancy-related outcomes. Groups such as older adults, people with
preexisting health conditions or multiple chronic diseases, and outdoor workers are
disproportionately susceptible to climate-associated health effects. Even as non-climate factors,
including adaptation measures, can help people cope with harmful impacts of climate change,
they cannot remove the risk of harm.
(4) Changes in climate resulting from human-caused emissions of greenhouse gases harm
the welfare of people in the United States. Climate-driven changes in temperature and
precipitation extremes and variability are leading to negative impacts on agricultural crops and
livestock, even as technological and other changes have increased agricultural production.
Climate change, including increases in climate variability and wildfires, is changing the
community composition and function of forest and grassland ecosystems and the services they
provide. Climate-related changes in water availability and quality vary across regions in the
United States with some regions showing a decline. Climate-related changes in the chemistry and
the heat content of the ocean are having negative effects on calcifying organisms and contributing
to increases in harmful algal blooms. U.S. energy systems, infrastructure, and many communities
are experiencing increasing stress and costs owing to the effects of climate change.
(5) Continued emissions of greenhouse gases from human activities will lead to more climate
changes in the United States, with the severity of expected change increasing with every ton
of greenhouse gases emitted. Despite successful efforts in many parts of the world to reduce
emissions, total global GHG emissions have continued to increase, and additional warming is
certain. All climate models—regardless of assumptions about future emissions scenarios or
estimates of climate sensitivity—consistently project continued warming in response to future
atmospheric GHG increases. Applying fundamental physics of the Earth system leads to the same
conclusion. Continued changes in the climate increase the likelihood of passing thresholds in
Earth systems that could trigger tipping points or other high-impact climate surprises.
In summary, the committee concludes that the evidence for current and future harm to human
health and welfare created by human-caused GHGs is beyond scientific dispute. Much of the
understanding of climate change that was uncertain or tentative in 2009 is now resolved and new threats
have been identified. These new threats and the areas of remaining uncertainty are under intensive
investigation by the scientific community. The United States faces a future in which climate-induced
harm continues to worsen and today’s extremes become tomorrow’s norms.
Prepublication Copy
1
Introduction
In 1896, Svante Arrhenius made a bold hypothesis: If gases that absorb heat energy are added to
Earth’s atmosphere, then science could quantify how much the average temperature of the Earth would
increase (Arrhenius, 1896). This hypothesis was based on earlier laboratory experiments showing that
carbon dioxide (CO2) and water molecules absorb energy, specifically at wavelengths typically emitted as
heat from the Earth’s surface (Foote, 1856; Tyndall, 1863). Gathering evidence to test the quantification
of planetary-scale effects of changing the composition of the atmosphere would be much more
challenging.
Climate records from weather stations and ship logs extend back to the 1700s, but it was not until
the late 1950s, during the International Geophysical Year, that multiple scientific disciplines came
together to more fully observe the Earth system (McCahey, 2025). The first continuous monitoring of
atmospheric CO2 began at this time (Keeling et al., 2001). Since then, an expanding array of increasingly
sophisticated evidence, spanning many aspects of the climate system and the natural environment, has
enabled the scientific community to test Arrhenius’ hypothesis.
An upward trend in atmospheric CO2 was documented by the early 1960s, confirming
expectations that human activities during the industrial era were changing the concentration of CO2 in the
atmosphere (Keeling et al., 2001). Nonetheless, recognizing the large variability inherent to the climate
system, scientists made sure that the long-term trend was robust before drawing conclusions about the
potential global impact of increases in CO2. For example, the National Research Council first addressed
the topic in a 1979 report, 20 years after atmospheric CO2 measurements were available (NRC, 1979).
These early climate change studies discussed the many uncertainties and unknowns that have
been the subject of research over the intervening years. At the same time, these reports called attention to
the potentially critical implications for people and the environment of changing climate and stressed the
importance of science to inform policy decisions. The foreword to the 1979 NRC report stated that its
conclusions might be “disturbing to policymakers,” noting that “a wait-and-see policy may mean waiting
until it is too late.”
This report provides an updated overview of the scientific evidence related to emissions of long-
lived greenhouse gases (GHGs) to the atmosphere, how the changing atmospheric composition is
affecting the climate system, and the impacts on human health and welfare. It is intended to inform
policymakers, and the public more generally, as they navigate many climate-sensitive decisions.
On August 1, 2025, the U.S. Environmental Protection Agency (EPA) issued a notice of proposed
rulemaking to rescind its prior findings that GHG emissions endanger human health and welfare (see Box
1.1) and invited public comments on its proposal (Reconsideration of the Endangerment Finding, 2025).
The “Endangerment Finding” was made by the EPA Administrator in 2009 (Endangerment Finding,
2009) and was informed by a Technical Support Document that reviewed scientific evidence available at
the time (EPA, 2009). Recognizing that significantly more is known today, the National Academies
launched this study to review newly available scientific evidence. To best inform EPA’s decision process,
the study was completed during the public comment period.
Prepublication Copy 3
The U.S. Supreme Court’s 2007 decision in Massachusetts v. Environmental Protection Agency (549 U.S.
497 (2007)) held that carbon dioxide (CO2) and other greenhouse gases (GHGs) fall within the statutory definition
of an “air pollutant” in the Clean Air Act. As a result, EPA was required to determine whether GHGs endanger
public health or welfare, whether emissions from mobile sources cause or contribute to that endangerment, and if
so, whether those emissions should be subject to regulation.
In 2009, EPA released “Endangerment and Cause or Contribute Findings for Greenhouse Gases Under
Section 202(a) of the Clean Air Act,” 74 FR 66496 (referred to as the Endangerment Finding). In this finding, the
EPA Administrator concluded that “six greenhouse gases taken in combination endanger both the public health
and the public welfare of current and future generations” and that “the combined emissions of these greenhouse
gases from new motor vehicles and new motor vehicle engines contribute to the greenhouse gas air pollution that
endangers public health and welfare under CAA section 202(a).” These conclusions were based on a review of
available scientific literature, summarized in a companion Technical Support Document (EPA, 2009). The
conclusions highlighted in the Executive Summary of EPA (2009) are included in Appendix C of this report.
The current study committee was charged to review the latest scientific evidence on whether
GHG emissions are reasonably anticipated to endanger public health and welfare in the United States (see
Box 1.2). The report focuses on evidence gathered by the scientific community since the publication of
EPA (2009) and describes supporting evidence, the level of confidence, and areas of disagreement or
unknowns.
This report does not address other factors that EPA considers in determining whether to regulate
emissions. The question raised in the proposed rulemaking of whether EPA has authority to regulate
GHGs to address climate change under section 202(a) of the Clean Air Act is outside the scope of the
report. The committee addresses the question of whether future emissions could cause or contribute to
future harm but does not consider specific scenarios in detail. Quantifying how emissions from new
motor vehicles or engines (as well as other non-vehicle sources) might cause or contribute to future GHG
emissions requires information—about future technology developments and regulatory scope—that is
unavailable to the committee. Addressing the feasibility of reducing emissions from motor vehicles and
other sources is outside the scope of this report but has been considered by other National Academies
studies (e.g., NASEM, 2021, 2024). Finally, the committee did not take up the question of whether
proposed regulations impose an undue economic burden. Such an assessment would require more detailed
information about proposed regulations and considerable further analysis of economic implications.
This fast-track study will review evidence for whether anthropogenic emissions of greenhouse gases to the
atmosphere are reasonably anticipated to endanger public health and welfare in the United States. The study will
focus on updates since the Environmental Protection Agency finalized the Endangerment Finding in 2009,
examine how current understanding compares to the 2009 Endangerment Finding, and provide explanation for any
changes. The study will develop conclusions that describe supporting evidence, the level of confidence, and areas
of disagreement or unknowns.
The causal chain from GHG emissions through to impacts on human health and public welfare is
shown in Figure 1.1. Assessing the impact of human-caused GHG emissions requires examining the
evidence for each step in this causal chain, as well as understanding the mechanisms that link the steps.
The report is organized to examine the evidence as follows:
Prepublication Copy
Introduction 5
• Chapter 2 examines observed changes in natural and human-caused GHG emissions, how
those emissions are changing the atmospheric concentration of GHGs, and the impact of
those changes on the energy imbalance at the surface of the Earth;
• Chapter 3 examines observed changes in climate conditions at the global and regional scales,
and the attribution of those climate changes to changes in human-caused GHG emissions;
• Chapter 4 addresses potential impacts of changes in GHGs on future climate, informed by the
evidence discussed in Chapters 2 and 3;
• Chapter 5 addresses the impacts of changes in GHGs and changing climate conditions on
human health; and
• Chapter 6 addresses the impacts of changing climate conditions on public welfare.
While this report focuses on the pathway by which changes in GHGs affect human health and
public welfare, many other non-climate factors also influence the impacts, as illustrated in Figure 1.1.
Non-climate factors can include changes in technologies or practices, changes in human systems (e.g.,
infrastructure, land use), changes in other environmental stressors (e.g., pollution), and other factors that
exacerbate or mitigate the impacts of human-caused GHG emissions. The committee discusses non-
climate factors, where significant, in examining impacts of changes in climate in Chapters 5 and 6.
FIGURE 1.1 Schematic of the causal chain from GHG emissions to impacts on human health and public welfare,
indicated by the darker blue arrows. The potential direct impacts of GHG emissions on human health (e.g., through
contribution to air pollution) and on ecosystems (e.g., changes in ocean chemistry) are indicated by the lighter blue
arrows. Non-climate factors, indicated by the grey circle and arrows, include changes in technologies or practices,
changes in human systems (e.g., infrastructure, land use), changes in other environmental stressors (e.g., pollution),
and other factors that exacerbate or mitigate the impacts of human-caused GHG emissions.
Prepublication Copy
This report summarizes the changes in evidence since 2009 for each step in the causal chain
shown in Figure 1.1, recognizing that the nature of the evidence and uncertainty varies across the different
links in this causal chain and for different impacts. The committee considered the science detailed in EPA
(2009) as the state of the science that informed EPA’s 2009 Endangerment Finding. Recognizing that
EPA (2009) covered a wide array of topics, the committee focused on those topics highlighted in the
Executive Summary of the document (see Appendix C). In addition, the committee identified other topics,
including those covered in the body of EPA (2009) but not highlighted in its Executive Summary, for
which new lines of discovery or impacts have emerged. The selection of topics to highlight was informed,
in part, by reviewing submissions to a public Request for Information (RFI).
In developing the conclusions of this report, the committee relied upon multiple sources of
evidence and considered conclusions to be stronger if there is broad agreement among independent lines
of evidence. The specific type of evidence relevant and available for each conclusion varies. The
committee weighed most heavily observational evidence, which includes direct measurements of
physical, chemical, or biological quantities; observations from space-based platforms and other
instruments that remotely sense the atmosphere, ocean, and biosphere; surveys of ecological variables;
and inventories and records of human systems (e.g., emissions data from industrial sources or
epidemiological studies). The committee also evaluated evidence from computational climate models that
simulate the Earth system. These models are an important line of evidence considered in this report. That
said, the climate system is complex and climate models are imperfect tools, therefore the committee relied
more heavily on observational evidence.
To prepare this report, the committee considered (1) widely available datasets that provide
information about GHG emissions, changes to the climate system, and human health and public welfare
impacts; (2) a broad range of peer-reviewed literature and scientific assessments; and (3) more than 200
comments submitted in response to a public RFI and through the standard feedback channels for National
Academies’ activities. In keeping with the study charge, the committee focused on literature published
since 2009 and on impacts to public health and welfare in the United States.
The committee examined scientific papers in the peer-reviewed literature, focusing on areas
where there have been significant new contributions that have changed or expanded understanding.
Where available, the committee drew on scientific assessments and National Academies’ reports that have
been prepared by large teams of scientists, incorporate mechanisms for broad public input, and are subject
to additional layers of peer review. These large-scale efforts provide periodic updates on the state of
knowledge. The committee tried to strike a balance between directly citing original studies and drawing
upon assessments, recognizing that both types of analyses provide useful information. Published
assessments, reports, and scientific papers provided useful input; however, it is important to note that the
committee then made its own determinations about how the evidence and understanding have changed
since 2009 in the key messages and conclusions of this report.
The EPA notice of proposed rulemaking cited a document, “A Critical Review of Impacts of
Greenhouse Gas Emissions on the U.S. Climate,” authored by a Climate Working Group assembled by
the Secretary of Energy. 1 This document was made available in draft form on July 29, 2025, at which
point the U.S. Department of Energy invited public comment and indicated it would be revised. Because
this document was not available in its final form and may change in response to comments received, the
committee does not cite it. Nonetheless, there is significant overlap in the topics addressed and in some of
the literature cited in the document and this committee’s report. For example, the committee addresses the
direct impacts of CO2 on the environment in a discussion of ocean heat and chemistry in Chapter 3 and in
a discussion of the drivers of ecosystem change in Chapter 6.
1
See https://www.federalregister.gov/documents/2025/08/01/2025-14519/notice-of-availability-a-critical-review-of-
impacts-of-greenhouse-gas-emissions-on-the-us-climate.
Prepublication Copy
Introduction 7
The climate varies significantly across the United States, spanning tropical to Arctic temperatures
and a wide range of precipitation regimes. Observed changes in the climate system similarly exhibit
regional variability. The factors that influence human health and welfare also vary across regions,
reflecting the differing geographies, ecosystems, infrastructure, economic activity, and recreation
activities found across the country. This report highlights selected regional differences in climate effects
and provides examples of regionally specific impacts, but it does not include a comprehensive assessment
of climate impacts in specific regions.
The committee focused on impacts to human health and welfare in the United States, similar to
the scope of EPA (2009). The Clean Air Act requires the EPA Administrator to take actions to safeguard
the American people. Thus, this report focuses on the risks to the U.S. population, who are most directly
affected by changes in climate conditions within its borders. In addition, the committee considered
changes to global oceans and the effects of these changes on U.S. coasts and fisheries.
EPA (2009) includes a section on impacts in other world regions, but the committee did not
address these impacts in detail. Even so, it is worth noting that changes to climate conditions in other
parts of the world do affect Americans. Many U.S. businesses are multinational, some with climate-
sensitive operations (e.g., supply chain agreements, shipping, agriculture). Many Americans live in other
parts of the world—in U.S. territories, as members of the armed forces and Foreign Service, and as part of
a large American expatriate population–and these places face their own climate risks. A growing body of
research has examined how climate stresses in other parts of the world can indirectly affect U.S. national
security (e.g., increasing migration pressures or creating political instability) (DOD, 2021; NASEM,
2023).
Climate affects human health and welfare in a multitude of ways, influencing everything from
how land is used for agriculture to how and where buildings are constructed to the diseases and other
health risks prevalent in different locations. As a result, the potential scope for the effects of GHGs on
welfare is particularly vast. The Clean Air Act identifies many effects that fall under “welfare” and leaves
open applications to other areas, as well. Section 302(h) of the Act (42 U.S.C. § 7602(h)) states: “All
language referring to effects on welfare includes, but is not limited to, effects on soils, water, crops,
vegetation, man-made materials, animals, wildlife, weather, visibility, and climate, damage to and
deterioration of property, and hazards to transportation, as well as effects on economic values and on
personal comfort and well-being.”
Given the very broad definition of welfare in the Clean Air Act and the broad scope covered in
EPA (2009), an exhaustive review of the rapidly growing body of literature on potential effects was not
feasible in this short report. In the discussion of impacts in Chapters 5 and 6, the committee chose to
focus on potential impacts with (1) more direct attribution to changes in climate conditions, and (2) more
direct impacts on human health and well-being. This choice is not intended to minimize the numerous
more complex or indirect ways that changing climate conditions affect people and nature. Rather it
reflects the committee’s determination that the evidence presented for this subset of impacts is sufficient
to support their conclusions regarding endangerment.
This report highlights examples of economic analyses but does not attempt an exhaustive review
of the growing body of literature related to economic impacts of GHGs or climate changes. Furthermore,
studies on future economic impacts are not included. Predicting potential future economic impacts
requires assumptions about many factors unrelated to climate that affect society and markets.
Similar to EPA (2009), this report does not consider the potential of adaptation measures to limit
future impacts on public health and welfare. In making a finding of endangerment, the Clean Air Act
requires only scientific determination of risk of harm, without speculation of potential actions that might
be taken to limit that harm. Predicting the efficacy of potential future adaptation actions to limit that risk
Prepublication Copy
requires assumptions about human behavior, government policies, technological advances, and many
other factors unrelated to climate.
Finally, the last century has been a time of rapid population growth, urbanization, increases in per
capita consumption, and technological innovation. As illustrated in Figure 1.1, this report describes a
range of factors that significantly affect human health and welfare, recognizing that the impacts of
changing climate occur in combination with other changing conditions. For example, trends in economic
damages from extreme weather events depend on the changes in the frequency, severity, or location of
extreme weather in combination with changes in where people live and the value of property located in
vulnerable places. This report draws on the large body of research examining impacts from climate
change in the context of other changes.
Prepublication Copy
2
Natural and Human-Caused Influences on
Earth’s Energy Imbalance
Emissions of greenhouse gases from human activities are increasing the concentration of
these gases in the atmosphere. Human activities, such as the extraction and burning of fossil fuels,
cement and chemical production, deforestation, and agricultural activities, emit greenhouse gases
(GHGs), which include carbon dioxide (CO2), methane (CH4), nitrous oxide (N2O), and fluorinated gases
(F-gases), to the atmosphere.
Total global greenhouse gas emissions continue to increase, even though U.S. emissions of
CO2 have decreased slightly in recent years largely due to changes in energy production and
consumption. The United States has the highest cumulative emissions 1 and among the highest per capita
emissions of GHGs in the world. The most recent decade (2010-2019) marked the largest decadal
increase in global CO2 emissions on record. Since 2009, methods of monitoring, observing, and
synthesizing inventories of these emissions have improved.
Increased greenhouse gases in the atmosphere are changing Earth’s energy balance, which
governs the physics and dynamics of the climate system. Earth’s climate is regulated by its radiative
energy balance—the difference between solar energy absorbed by the surface and atmosphere and
infrared energy emitted back to space. GHGs absorb and re-emit infrared radiation (energy) in all
directions, including downward to the surface and upwards to space. The overall imbalance (net total of
human and natural forcings 2) has led to warming of the climate system since 1750. Advances in remote
sensing, longer term records, and analyses provide continued and more robust support for conclusions
about Earth’s energy imbalance than was available in 2009.
Multiple lines of evidence show that greenhouse gas emissions from human activities are the
primary driver of the observed long-term warming trend and other changes in Earth’s energy
balance, and that natural forces cannot account for observed changes. No known natural drivers,
such as incoming solar radiation or volcanic emissions, can explain observed changes. This is particularly
true for the magnitude of warming at Earth’s surface and the vertical distribution of warming in the
troposphere (lower atmosphere) and cooling in the stratosphere (upper atmosphere). These changes are
consistent with the physics and dynamics of the climate response to GHG increases.
Statements from EPA (2009) continue to be true and are supported by improved scientific
evidence. Evidence for human-caused increases in greenhouse gas emissions, their effects on Earth’s
climate, and the attribution of climate change to human activities has grown stronger and more
compelling, backed by multiple, independent lines of improved data and analysis.
EPA (2009) considered six well-mixed GHGs emitted by human activities: CO2, CH4, N2O,
HFCs, PFCs, and SF6 (Box 2.1). Both global and U.S. emissions of GHGs are dominated by those of
1
Cumulative emissions are the total sum of GHGs released over time. This is usually calculated using 1750, when
the industrial revolution began, as a starting point.
2
Forcings are factors that influence Earth’s radiative balance. Human forcings (or “anthropogenic” forcings) consist
of the emissions of GHGs and land-use change. Natural forcings are nonhuman factors that affect Earth’s radiative
balance. This includes factors such as solar radiation, Earth’s orbital cycle, and volcanic activity.
Prepublication Copy 9
CO2. Emissions of other GHGs are often reported as the amount of CO2 emissions that would produce an
equivalent amount 3 of radiative forcing (the imbalance between incoming and outgoing energy in Earth’s
atmosphere; see Section 2.3) over a 100-year period, which is referred to as carbon dioxide equivalents,
or “CO2e.”
BOX 2.1 Legislative Requirements Applying to Greenhouse Gas (GHG) Emissions Regulations
The 2009 Endangerment Finding (see Box 1.1) identified carbon dioxide (CO2), methane (CH4), nitrous oxide
(N2O), hydrofluorocarbons (HFCs), perfluorocarbons (PFCs), and sulfur hexafluoride (SF6) as the primary GHGs
emitted by human activities not already covered by the Montreal Protocol (which controls additional GHGs:
chlorofluorocarbons [CFCs] and hydrochlorofluorocarbons [HCFCs]). Three of the GHGs identified by the
Endangerment Finding—CO2, CH4, and N2O—are vehicle as well as stationary source emissions. In addition to
the GHGs identified in the Endangerment Finding, vehicle emissions also include “criteria pollutants”—nitrogen
oxides, carbon monoxide, sulfur oxides and particulate matter—which are associated with direct human health
effects.
As recognized in the Endangerment Finding, the Clean Air Act Section 202(a) requires EPA to create
emission standards to limit air pollutants that endanger public health or welfare. These regulations focused
initially on criteria pollutants but apply also to vehicle GHG emissions. Additionally, the Clean Air Act broadly
together with the American Innovation and Manufacturing Act of 2020 require EPA to control, from stationary
sources, all GHGs included in the Endangerment Finding.
At the time of the Endangerment Finding, EPA vehicle exhaust emission standards regulating criteria
pollutants had been in place for many years. Vehicle CH4 and N2O emissions have been subsequently regulated
since 2015. Vehicle emission standards for CO2 are planned to start in 2027. CO2 emissions have also been
regulated indirectly by evolving Corporate Average Fuel Economy standards, administered by the National
Highway Traffic Safety Administration since 1975. CO2 from stationary power plants has recently been regulated
by EPA. Two of the gases identified in the Endangerment Finding, HFCs and PFCs, are regulated with stationary
source emissions limits, and SF6 is limited in some cases and required to be reported in others.
In its 2022 assessment, the Intergovernmental Panel on Climate Change (IPCC, Dhakal et al.,
2022) estimated that total global annual emissions of GHGs had increased to approximately 59 billion
metric tons CO2e per year (gigatons per year, Gt/y). During the decade beginning in 2010, the average
emissions were 56 GT/y, which is 9 Gt/y (or 16%) higher than the decade from 2000-2009, leading to the
greatest decadal growth in atmospheric concentrations in the modern record. In contrast, over the same
period GHG emissions from the United States have decreased (Figure 2.1). EPA (2009) used the latest
data available at the time (from 2007) to estimate U.S. emissions as 7.150 Gt/year of CO2e. Since that
time, improvements in the emission estimation procedures used by the EPA facilitated refining the
estimate of 2007 emissions to 7.530 Gt/y CO2e (EPA, 2024a). Estimated emissions for 2022, the latest
year for which EPA has developed a comprehensive national emission estimate, are 6.343 Gt/y CO2e.
This decrease in total emissions is largely the result of changes in the relative amounts of different energy
sources used in the United States.
Since 2007, the United States has transformed from an energy importing country to the world’s
largest producer of oil and natural gas and a major exporter of energy. Major trends with consequences
for GHG emissions include widespread substitution of coal used for electricity generation with natural gas
and renewable sources of energy and increases in the efficiency of petroleum use for transportation
(Figure 2.2). The changes in electricity generation have led to decreases in CO2 emissions, because
3
Global Warming Potential is a common metric in discussions of GHGs used in EPA (2009). Global Warming
Potential compares how much total heat a greenhouse gas traps over a chosen time horizon (e.g., 100 years) relative
to CO₂. Carbon dioxide equivalents (CO₂e) express a GHG quantity as the amount of CO₂ that would cause the same
trapping of heat over that horizon, calculated as mass multiplied by Global Warming Potential.
Prepublication Copy
burning coal results in higher emissions of CO2 per unit of energy produced than burning natural gas or
use of renewable sources of energy.
Improvements in the fuel efficiency of vehicles and introduction of electric vehicles has led to a
relatively constant to slightly decreasing amount of petroleum use, despite increases in total miles
travelled by vehicles in the United States (U.S. Bureau of Transportation Statistics, 2025). While total
emissions in the United States have decreased due to these changes in patterns of energy use, the country
still remains one of the largest sources of GHG emissions both in total and per capita, and the largest
cumulative emitter of GHGs (IPCC, 2022; Jones et al., 2024a).
FIGURE 2.1 Emissions of CO2, CH4, N2O, and several fluorinated gases (HFCs, PFCs, SF6, and NF3) in the United
States from 1990 to 2022. Data: EPA, 2024a.
SOURCE: EPA, https://www.epa.gov/climate-indicators/climate-change-indicators-us-greenhouse-gas-emissions.
Accessed September 2025.
Changes in energy use have also affected emissions other than GHGs. EPA (2009) noted that
emissions of sulfur dioxide (S2O) and particulate matter (particles less than 10 microns in diameter,
PM10), which accompany the burning of fossil fuels, had decreased in the decades prior to 2009. S2O
emissions have continued to decrease since 2009 due to shifts away from the use of coal in electricity
generation and the lowering of the sulfur content in fuels in diesel fueled vehicles. As reported through
EPA’s National Emissions Inventory, 4 emissions of S2O decreased from 16.3 million tons per year in
2000 to 8.0 million tons per year in 2009 (EPA, 2025a). From 2010 to 2020, total emissions were further
reduced to 1.8 million tons per year. These decreasing emissions in S2O led to reduced formation of
sulfates found in particulate matter in the atmosphere. Because sulfates in particles have a net cooling
effect, the reduced sulfate concentrations would be expected to lead to reduced cooling.
4
See https://www.epa.gov/air-emissions-inventories/national-emissions-inventory-nei.
Prepublication Copy
FIGURE 2.2 U.S. energy production 1990 to 2025 (top) and consumption (bottom). The production and use of
energy in the United States have evolved since 2007, as tracked by the U.S. Department of Energy’s Energy
Information Administration.
SOURCES: U.S. Energy Information Administration, 2025a, 2025b.
Direct emissions of particulate matter, not including those associated with wildfires, are more
complex, having decreased by approximately one third from 2000 to 2009 then remaining relatively
constant since. The quantity of particulate matter emissions from wildfires varies widely from year to
year. For example, in 2022, a year which had large areas burned by wildfires, PM10 emissions from
wildfires were estimated to be an order of magnitude larger than in 2009. The wildfire emissions in 2022
Prepublication Copy
occurred in localized areas for limited time periods but were estimated to increase total national PM10
emissions for the year by roughly 20%, resulting in very high concentrations of PM10 during the time
periods when the fires were occurring (EPA, 2025h).
Estimates of emissions of GHGs and other pollutants have associated uncertainty from available
data and methodologies. Among GHGs, estimates of CO2 emissions have the lowest uncertainties,
because the majority of these estimates are based on fuel consumption data, which are accurately and
precisely tracked, multiplied by the emissions per usage, which is also well known. Estimates of
emissions of other GHGs, such as CH4, have higher uncertainties, but these uncertainties have been
reduced by advances in measurement technologies over the last decade (NASEM, 2018). In reviewing the
status of CH4 emission measurements, the National Petroleum Council (2024) concluded in a report to the
U.S. Secretary of Energy that “new measurement technologies have emerged over the last five years that
are dramatically improving emission detection and the accuracy of emission estimates.”
Observations of atmospheric concentrations provide additional evidence of increasing global
GHG emissions. Concentrations of GHGs in the lower atmosphere, such as those measured at the Mauna
Loa monitoring station in Hawaii, are increasing (Figure 2.3). Figure 2.4 shows annual increases in N2O
concentrations, which have been increasing at a generally accelerating rate, and CH4 concentrations, for
which emission increases have had more complex behavior but overall have been increasing at a
generally accelerating rate since 2009.
FIGURE 2.3 Atmospheric CO2 measurements and trends from Mauna Loa Observatory. These graphs compare the
rise of atmospheric CO2 measured at Mauna Loa (Left: monthly mean 1960-2025; Right [bottom]: annual average
increase) and global records (Right [top]: global annual increase from the Global Greenhouse Gas Reference
Network 5). The decadal average rate of increase of CO2 in the graphs on the right is depicted by the black,
horizontal lines. Rates of increase measured at Mauna Loa align closely with global records.
SOURCE: NOAA Global Monitoring Laboratory, https://www.noaa.gov/news-release/during-year-of-extremes-
carbon-dioxide-levels-surge-faster-than-ever.
5
The Global Greenhouse Gas Reference Network is based in NOAA’s Global Monitoring Laboratory and uses
observations from observatories, aircore samples from ballons, and air samples taken by small aircrafts and
volunteers.
Prepublication Copy
FIGURE 2.4 Annual increases in atmospheric nitrous oxide (N2O; top) and methane (CH4; bottom) based on
globally averaged marine surface data. The baseline annual atmospheric concentration (0 line) of N2O in 2001 is
316.36 parts per billion (ppb) (top) and of CH4 in 1984 is 1644.84 ppb (bottom).
SOURCE: Lan and Dlugokencky, 2022, accessed via NOAA at https://gml.noaa.gov/ccgg/trends_n2o and
https://gml.noaa.gov/ccgg/trends_ch4.
The most recent average decadal increase in CO2 concentration is more than 2 parts per million
(ppm) per year. This rate of increase is more than 100 times faster than natural increases, such as those
that occurred at the end of the last Ice Age 11,000-17,000 years ago. Further documentation of increases
in atmospheric GHG emissions comes from the relative abundance of carbon isotopes in the observations.
Measured decreases in the fraction of the carbon isotopes 14C and 13C in ice core records show that the
rise in CO2 is largely from combustion of fossil fuels, which have low 13C fractions and no 14C (NRC,
2020). Figure 2.5 shows these longer historic trends combined with observations from recent atmospheric
concentrations, which demonstrate the steep rate of increase beginning in approximately 1750. Beyond
Prepublication Copy
these historical records, fundamental physical and chemical processes also indicate that once emitted,
CO2 persists in the atmosphere for centuries, 6 ensuring today’s emissions will contribute to atmospheric
concentrations far into the future.
FIGURE 2.5 Global concentrations of the primary human-caused greenhouse gases over the last approximately
2000 years (blue dotted lines), with the shaded areas (orange) indicating the time period since EPA (2009) was
published. The timescale (x-axis) for left panel spans 2000 years; the timescale for right spans only the past 65
years.
SOURCES: Data in the left panels are from ice cores (Ahn, 2023) while data in the right panels are from NOAA
observations (Lan and Dlugokencky, 2022, accessed September 2025; NOAA GML, 2025).
6
For more information about the global carbon cycle, see “Global Carbon Budget 2024” (Friedlingstein et
al., 2025).
Prepublication Copy
Understanding how energy flows through the Earth system, quantifying those flows in the form
of energy fluxes, and evaluating the factors that influence and change these fluxes is the foundation for
understanding the physical climate system and is the basis for characterizing and predicting changes to
Earth’s climate. Changes in energy flows, referred to as “forcings,” operate on a variety of timescales and
can be human-caused or from natural sources. The influence of any individual forcing is typically compared
using a measure of its impact on the Earth’s radiative balance, measured as the annual average net change in
radiation (in Watts) at the top of the atmosphere per square meter of the planet’s surface (W/m2).
Since 2009, advances in observations and the analysis of expanding time series have shown that
Earth has a positive energy imbalance —i.e., more energy is coming in than going out—and this
imbalance is increasing over time (Hakuba et al., 2024; Loeb et al., 2021). Over the longer-term, this
imbalance will drive additional net heating of the planet. Human-caused, or anthropogenic, forcings are
the primary driver of the imbalance in Earth’s energy and resulting heating. Total anthropogenic forcing
comes mainly from GHG emissions from human activities but also includes contributions from aerosols
(particulates that are mostly cooling agents) and changes in the Earth’s surface cover. Combining all of
these human-caused contributions leads to a best estimate of +2.97 (+2.05 to +3.77) W/m2 in net total
anthropogenic radiative forcing for 1750 to 2024 (Forster et al., 2025). This estimate has increased by
more than 80% since the estimate cited in EPA (2024); the net total anthropogenic forcing for 1750 to
2005 was +1.6 W/m2 (+0.6 to +2.4 W/m2).
The international community has developed a robust and comprehensive understanding of Earth’s
energy balance, along with its associated uncertainties, through decades of careful and sustained
assessments (e.g., Stephens et al., 2022). Today, more objective methodologies that jointly constrain both
the water and energy budgets are used (L’Ecuyer et al., 2015; Rodell et al., 2015), moving away from
prior, more ad hoc, methods used at the time of EPA (2009). These methodologies support the finding
that Earth, in the annual mean, currently experiences a radiative imbalance at the top of the atmosphere,
referred to as Earth’s Energy Imbalance (EEI). While net radiative forcing represents the combined effect
of external drivers of Earth’s energy balance, EEI is the portion of the forcing not yet offset by the
planet’s radiative response and observed directly as the annual mean of Earth’s heat uptake. EEI is
measured as the sum of all heat content changes occurring in the ocean, land, atmosphere, and cryosphere.
EEI is the most holistic picture to date of heat accumulation by the Earth system and a quantitative
measure of the cumulative impact of both natural and anthropogenic radiative forcings and feedbacks.
Independent estimates of the heat uptake, taking account of all heat content changes, largely match
estimates of EEI (Hakuba et al., 2024; Meyssignac et al., 2019; von Schuckmann et al., 2016; see Figure
2.6).
Approximately 90% of Earth’s heat uptake occurs in the ocean. Hakuba et al. (2024) offer a
comprehensive review of 18 different sources of ocean heat content data and conclude that the range of
annual mean EEI falls between 0.40 to 0.96 W/m2 where the spread is due to differing sources of ocean
data, mapping methods, and quality control procedures across ocean data products (Figure 2.6). They also
assess the rate of change of EEI over the observation record and, while ranging from 0.1 to 1.0 W/m2 per
decade, conclude that all major sources of data indicate an increasing rate of EEI over the past
approximately 20 years. This pattern points to an accelerating warming of the planet (e.g., Mauritsen et
al., 2025), which is also consistent with observations that suggest an accelerated rise in sea level due in
part to thermal expansion of sea water (see also Section 3.5).
EEI methods and estimates provide more robust support for the EPA (2009) conclusion that “the
global average net effect of the increase in atmospheric GHG concentrations, plus other human activities
(e.g., land-use change and aerosol emissions), on the global energy balance since 1750 has been one of
warming.”
Prepublication Copy
FIGURE 2.6 The Monitoring Ocean Heat Content and earth energy imbalANce (MOHeaCAN) (Legos) time series
of ocean heat uptake (OHU, blue) and the 90% confidence interval of the Clouds and the Earth’s Radiant Energy
System Energy Balanced and Filled (CERES EBAF) net radiative flux (EEI, black). Both time series are low-pass
filtered at 3 years cut-off time (Lanczos) which removes high-frequency noise related to intrinsic ocean variability.
Legos ocean data and CERES Top of Atmosphere (TOA) radiative flux data broadly agree that the Ocean Heat
Uptake is increasing over time.
SOURCE: Hakuba et al., 2024. CC BY 4.0.
Over timescales of decades to centuries, the most important known natural forcings are solar
output and volcanic eruptions. Prior to modern measurement techniques, estimates of long-term changes
in the output of the Sun have substantial uncertainties. Nevertheless, solar forcing from the preindustrial
average over the ~11-year solar cycle to the average over the last complete solar cycle (2009-2019) is
estimated by the IPCC’s Sixth Assessment Report and more recent analyses (Forster et al., 2025) to be
+0.01 W/m2 (−0.06 to +0.08; 90% 7 confidence interval). The best estimate of solar forcing is roughly 300
times less than anthropogenic forcing. Even the high end of the range of solar forcing is equal to only a
few percent of anthropogenic forcing over this period. Uncertainties in observations since the
preindustrial period do not support conclusions about trends in solar forcing over that period. However,
over the last approx. 45 years, during which satellite observations are available (leading to higher
confidence in the observed trends), it is very likely that solar forcing has decreased (Amdur and Huybers,
2025; Matthes et al., 2017; Montillet et al., 2022). This likely decrease in solar forcing was observed at
the same time that the Earth has been warming at its most rapid pace since the preindustrial period.
Volcanic forcing is highly irregular and sporadic, with large eruptions which inject material into
the stratosphere driving short-term cooling but with no evidence for long-term trends over the last two
centuries (Forster et al., 2021). CO2 emissions from volcanoes are negligibly small. Hence, natural
forcing is both very small over the time since industrialization and has very likely caused a minor amount
of cooling over recent decades rather than contributing to the observed warming.
7
The range in parentheses included with radiative forcing estimates represents the “very likely” range, meaning
there is a 90% likelihood that the actual value falls within this range.
Prepublication Copy
Owing to the successful implementation of the Montreal Protocol (1987 international treaty to
control ozone-depleting substances), concentrations of several fluorinated GHGs (CFCs) have decreased
whereas concentrations of other fluorinated gases are still increasing (i.e., CFC-replacements, such as
hydrofluorocarbons and the industrial PFCs; Forster et al., 2025). The 2024 estimate of radiative forcing
relative to the 1750 baseline 8 from halogenated gases (primarily fluorine-containing, or F-gases) is +0.41
(+0.33 to +0.49) W/m2 (Forster et al., 2025), with relatively flat growth rates owing to the offsetting
influence of declining CFCs and increases in other F-gases. These gases are entirely anthropogenic in
origin.
CO2 concentrations have increased by ~50% (through 2023) relative to the mean of the range
seen over 1000 to 1850, with analogous increases of ~160% and 25% for CH4 and N2O, respectively
(Figure 2.5). The observed concentration increase in these gases is attributable almost exclusively to
anthropogenic activities based on analyses of budgets for each GHG and on the isotopic signature of
atmospheric carbon (see Figure 2.7). Current CO2 levels are likely the highest they have been in the last 3
million years (Canadell et al., 2021; de la Vega et al., 2020; Martínez-Botí et al., 2015), and the rates of
change in all three of the main GHGs (CO2, CH4, and N2O) are faster than those seen at any time in the
last million years of ice core records (Canadell et al., 2021).
Understanding of the effect of GHGs on the Earth’s energy balance remains solidly grounded in
physics and in laboratory measurements, which date back to the 19th century (e.g., Tyndall, 1863), as
well as in surface and satellite measurements (e.g., Harries et al., 2001; Teixeira et al., 2024). This
understanding of the fundamental physics of the Earth’s energy system, combined with observational
constraints on feedback 9 processes, mean that, at the global scale, the effect of a GHG forcing can be
evaluated with a simple equation and does not require the use of complex numerical models or other
complicated analysis. In practice, relatively simple models used decades ago can now be seen to have
performed extremely well in matching the observed global mean warming over time per unit radiative
forcing (Hausfather et al., 2020; Supran et al., 2023).
Since EPA (2009), the radiative forcing due to increasing atmospheric concentrations of the three
main human-caused GHGs has increased from the +2.30 (+2.07 to +2.53) W/m2 estimate in 2005 to +3.13
(+2.7 to +3.6) W/m2 in 2024 (Forster et al., 2025). This rapid rise in forcing continues the trend reported
in EPA (2009) that “the rate of increase in positive radiative forcing due to these three GHGs during the
industrial era is very likely to have been unprecedented in more than 10,000 years.”
Though not emitted directly by anthropogenic activities, ozone is another important GHG. Ozone
in the lower atmosphere (the troposphere) has a large warming impact on climate, especially in the upper
troposphere. Ozone in the upper atmosphere (the stratosphere) has only a weak effect on climate but is
important in protecting the surface from ultraviolet radiation. Tropospheric ozone has increased since
industrialization due to human-caused emissions of ozone precursor gases that lead to photochemical
ozone production. The effect of this tropospheric ozone increase has outweighed the impact of ozone loss
in the stratosphere, leading to a net positive forcing of +0.50 (+0.25 to +0.75) W/m2 in 2024 (Forster et
al., 2025). Importantly, roughly half the radiative forcing from tropospheric ozone increases is due to
emissions of CH4, which contributes to the chemical formation of ozone in the troposphere. Hence, the
overall climate influence of the three main GHGs is larger than that attributed to changes in their
concentrations alone.
8
Unless otherwise specified, estimates of radiative forcing (W/m2) are given relative to a preindustrial (1750)
baseline.
9
Feedbacks are natural or human-induced processes that impact a system (i.e., hydrologic, climate, atmospheric).
Positive feedback amplifies the initial change, and negative feedback decreases the initial change.
Prepublication Copy
Aerosols affect radiative forcing both through direct influences on radiation and indirect effects
on cloud microphysics and lifetime, which impact the radiative properties of clouds. The IPCC’s Sixth
Assessment Report (IPCC, 2021) provides an updated assessment of this combined aerosol forcing.
Forster et al. (2021) conclude with high confidence 10 that the radiative forcing from aerosols is negative
(cooling) and estimate, with medium confidence, that for 2019 it is −1.1 (−1.7 to −0.4) W/m2; the total
effective radiative forcing from aerosols in 2024 is estimated to be −1.07 (−1.90 to −0.43) W/m2 (Forster
et al., 2025). This aerosol forcing (about −1 W/m2) does affect the energy balance, although the cooling
effect is only about a quarter of the warming associated with GHGs (about +4 W/m2 from combined CO2,
CH4, N2O, ozone, and F-gases forcings included above) over the Industrial Era.
As stated in EPA (2009), “[GHGs], once emitted, can remain in the atmosphere for decades to
centuries, meaning that (1) their concentrations become well-mixed throughout the global atmosphere
regardless of emission origin, and (2) their effects on climate are long lasting.” Atmospheric
concentrations of the three main GHGs—CO2, CH4, and N2O—have continued to increase since 2009 as
discussed in the previous sections, and it is virtually certain that this increase is due to human activities.
These resulting concentration increases in the atmosphere represent human-forcing on Earth’s energy
balance. The increased concentrations of well-mixed GHGs impact the climate though the increased
infrared radiation they emit. This enhanced emission downward to the surface is an expression of the
GHG radiative forcing and drives the surface and lower atmospheric warming. The enhanced emission of
infrared radiation upward to space, by contrast, results in a cooling of the upper atmosphere. This
warming-cooling dipole pattern, a unique signature of GHG radiative forcing of Earth’s climate, is well
documented in observations of temperature change (Santer et al., 2023).
No known natural process could account for observed increases in GHGs that both (1) are well
beyond the range of values seen over the past million years and (2) occur simultaneously for all three of
the main well-mixed GHGs. Furthermore, the isotopic record of CO2 (Figure 2.7, top) shows an increase
in very old, plant-based CO2 (depleted in 13C) that is a signature of fossil fuel burning (Graven et al.,
2020). Consistent with the attribution of increasing GHG concentrations to human activities, and based on
understanding of the radiative forcing of different climate drivers and the strength and speed of the
climate response to forcing, the best estimate of the anthropogenic contribution to the observed surface
warming of 2.23°F (1.24°C; 2 to 2.43°F / 1.11 to 1.35°C) for 2015-2024, relative to 1850-1900, is 100%
(Eyring et al., 2021; Forster et al., 2025; Gulev et al., 2021). Natural forcings are estimated to have
contributed only about 0.09°F (0.05°C; −0.18 to 0.36°F / −0.1 to 0.2°C) to this warming (Eyring et al.,
2021), which is higher than the contribution typically found when using the standard practice of
comparing periods that both occur at solar minima to avoid the decadal fluctuations of the ~11-year solar
cycle.
The observed vertical pattern of warming (lower atmospheric warming [Figure 2.7, middle],
upper atmospheric cooling [Figure 2.7, bottom]) is consistent with the effect of increasing GHGs but is
inconsistent with the effect of increased solar forcing (Casas et al., 2023; Santer et al., 2023). Thus, it is
virtually certain that observed warming is due to human activities (Figure 2.7).
10
When referencing IPCC assessments, statements that include likelihood or confidence levels follow the same
established scales used by those authors—e.g., for likelihood: “virtually certain” equates to 99–100% probability;
very likely 90–100%; likely 66–100%, etc., and for confidence: “very high confidence” refers to at least a 9 out of
10 chance of being correct and “high confidence” to at least 8 out of 10. For more information on these designations,
see IPCC, 2023.
Prepublication Copy
FIGURE 2.7 (Top) Concentration of CO2 (orange dots, increasing) alongside the fraction of CO2 containing the 13C
isotope (blue dots, decreasing) over the last 1,000 years. Total CO2 data in the upper panel is from Ahn (2023) and
NOAA observations (NOAA GML, 2025) with 13C data from Rubino et al (2019). (Middle and Bottom) Observed
(thick black lines) and modeled trends (other lines) in atmospheric temperatures in the middle troposphere (Middle)
and in the stratosphere (Bottom). The lower panels are reprinted with permission from Casas et al. (2023), with
observations from the Microwave Sounding Unit (MSU) updated from Mears and Wentz (2016) and observations
from the Stratospheric Sounding Unit (SSU) updated from Zou and Qian (2016).
Prepublication Copy
3
Observed Climate Changes from Human-Caused
Greenhouse Gas Emissions
Improved observations confirm unequivocally that greenhouse gas emissions are warming
Earth’s surface and changing Earth’s climate. Longer records, improved and more robust
observational networks, and analytical and methodological advances have strengthened detection of
observed changes and their attribution to elevated greenhouse gases (GHGs). EPA (2009) provided
evidence for a range of observed changes in Earth’s climate associated with elevated global
concentrations of GHGs. Many of the global trends could also be observed in the United States.
Observations show continuing increases in hot extremes alongside declines in cold extremes,
furthering the conclusion in EPA (2009). Six decades of observations document a tripling of average
annual heat-wave frequency since the 1960s. Heat metrics in many regions show heat-related changes.
For example, occurrence of the hottest day, warmest night, warm spells, and other heat events have
intensified in the southeast.
In the United States, regional shifts in annual precipitation and a higher number of extreme
single-day precipitation events have been observed. The amount of land area experiencing greater than
normal annual precipitation totals has increased since 1895, and the prevalence of extreme single-day
precipitation events has risen substantially since the 1980s. Regionally, the Northeast has experienced
about a 60% increase in the amount of precipitation falling on the heaviest 1% of days since 1958, with
the Midwest up roughly 45% over similar periods.
Observations show continued warming of the Earth’s oceans. An increase in global sea
surface temperature has been observed since 1900 and ocean heat content increases in the upper 2,000
meters are also evident since 1960 throughout the global oceans. Ocean warming has contributed to
increases in rainfall intensity, rising sea levels due to thermal expansion, the destruction of coral reefs,
declining ocean oxygen levels, and declines in ice sheets, glaciers, and ice caps in the polar regions.
Ocean pH has decreased, and along with ocean warming, poses risks to marine ecosystems
and the benefits they provide. Ocean pH decreased from 8.2 in 1750 to 8.1 today, consistent with the
finding in EPA (2009), and represents about a 30% increase in the hydrogen ion concentration in ocean
water (which equates to less alkaline and more acidic conditions). Changes in pH in U.S. offshore waters
track with the global average trends, but changes in U.S. coastal waters vary due to local conditions.
Decreasing ocean pH, along with warming, poses risks to species with shells and skeletons and coral reefs.
Global mean sea level has risen about 7 inches (approximately 18 centimeters) since 1900,
and the rate of sea-level rise is accelerating. Regional relative sea level rose on average by
approximately 11 inches in the last century along the continental United States, putting many coastal
communities at risk of increased coastal flooding and vulnerability to coastal storms. Changes in average
sea level have doubled the frequency of high tide flooding in the continental United States over the past
few decades.
Evidence of increasing wildfire severity linked to climate change has grown since EPA
(2009). Changing climate conditions, including warmer springs, prolonged summer dry periods, and drier
soils and fuel sources, have increased the likelihood for wildfire ignition and spread. The wildfire area
burned in the western United States has increased in recent decades, resulting in substantial increases in
fine particulate matter and other air pollutants.
The following sections provide more detail on the observed changes of key Earth system
components—temperature, precipitation and drought, oceans, cryosphere, and biosphere—since EPA
Prepublication Copy 21
(2009), including attention to trends, extremes, and regional variation across the United States, as well as
signals of human attribution to these observations.
3.2 TEMPERATURE
The Earth energy imbalance discussed in Chapter 2 leads to warming of the surface and lower
atmosphere, which is clearly detected in temperature observations. Global mean surface temperatures
have increased by 2.23°F (1.24°C; range 1 2.00 to 2.43°F / 1.11 to 1.35°C) for approximately the last
decade (2015–2024) relative 2 average (Forster et al., 2025). This temperature increase is approximately
60% greater than the warming reported in EPA (2009), reflecting the very rapid warming of the planet
during the last two decades. The warming rate reached a value of 0.49°F (0.27°C; 0.36 to 0.72°F / 0.2 to
0.4°C) per decade during 2015-2024, helping drive warming rates over the past 50 years and decadal
average temperatures both to their highest values in at least the last 2000 years (Forster et al., 2025; Gulev
et al., 2021).
Consistent with global trends, temperatures in the United States have increased and at an
increasing rate (Figure 3.1). Because the land warms more than the ocean, the U.S. annual mean has
increased more than the global mean. Since 1970, annual mean temperatures in the contiguous United
States have increased by 2.5°F (1.4°C) and by 4.2°F (2.3°C) in Alaska (Marvel et al., 2023) compared to
the global mean of approximately 1.7°F (0.9°C). This is a marked increase from the 1.3°F (0.7°C) U.S.
temperature increase reported in EPA (2009) for the 20th century, continuing the trend of the rate of
warming increasing, in addition to the actual temperature increase, over the past (now more than four)
decades.
Warming has been observed nationwide but is the most pronounced in Alaska, the West, and the
Northeast. The Southeast has historically experienced slower warming (a “warming hole” where long-
term temperature trends are negative or non-significant compared to the rest of the country), but this trend
has diminished in recent decades. The United States has also experienced longer summers and shorter
winters, with winters warming nearly twice as fast as summer in many northern states. Night temperatures
have also increased consistently in both summer and winter, which affects plants and ecosystems (see
Chapter 6). Temperature trends are observed and supported by instrumental surface networks and
homogenized datasets; homogenization 3 has improved and high-resolution reanalyses have sharpened the
detection of regional trends since 2009 (Eyring et al., 2021; Marvel et al., 2023).
Extreme Temperatures
Observations show continuing increases in hot extremes alongside declines in cold extremes since
2009, furthering the conclusion in EPA (2009) that “widespread changes in extreme temperatures have
been observed in the last 50 years across all world regions, including the United States.” Multiple
independent datasets concur that the frequency and intensity of record heat—hot days, hot nights, heat
waves—have risen while record cold—cold days, cold nights, and frost—have diminished over most land
areas across the globe, including the United States (Fischer et al., 2025). These findings have been
documented and reconfirmed in national and international climate assessments since 2009 (Eyring et al.,
2021; USGCRP, 2023), as well as by records compiled by EPA’s Climate Change Indicators (EPA,
2025f) for the United States.
1
Temperature ranges provided in parentheses represent the “very likely” or 90% confidence interval.
2
Many estimates in Chapter 2 used 1750 as a baseline; however, quantities provided in this chapter vary in baseline
depending on what characteristic is being measured and the start of long-term observational records for that
characteristic. These baselines are provided as appropriate for different observations.
3
Homogenized datasets adjust for non-climatic influences; for example, station moves or instrument changes, to
ensure observed trends reflect real changes rather than artifacts of data collection methods.
Prepublication Copy
FIGURE 3.1 (Top) Temperature in the United States. (Bottom) Rate of temperature change.
SOURCE: (Top) EPA, 2025f, https://www.epa.gov/climate-indicators/climate-change-indicators-us-and-global-
temperature. (Bottom) USGCRP, 2023.
Prepublication Copy
EPA’s Heat Waves indicator (EPA, 2025c), based on six decades of observations, documents a
tripling of average annual heat-wave frequency since the 1960s, with earlier starts and later ends to the
season; regional studies reinforce these patterns. Weather events are classified as “extreme” relative to
local historical baselines, so regional variability in these observations and events is expected. The
frequency of extreme heat events is significantly increasing in the western United States, while
seasonally-relative extreme heat events are increasing in parts of the South (Ibebuchi et al., 2025). The
Southeast experienced intensification of extremes—annual occurrence of the hottest day, the warmest
night, warm days, warm nights, summer days, tropical nights, and warm spells—over 1978-2017 (Fall et
al., 2021). Florida’s observed heat-stress metrics (heat index, wet-bulb globe temperature) have risen
markedly since 1950 (McAllister et al., 2022), while relative extreme cold events are decreasing the most
in the Florida peninsula, as well as in southern California and Nevada (Ibebuchi et al., 2025). Analyses of
cold air outbreaks and cold waves find broad declines in frequency, duration, and severity across the
northern mid-latitudes, including in the United States (Smith and Sheridan, 2020; van Oldenborgh et al.,
2019).
Extreme weather events most closely related to temperature have been found to be more frequent
and intense due to human-caused climate change using extreme event attribution science (NASEM,
2016). Non-climate factors can also drive extreme temperature events and have done so in the past in the
United States. The Dust Bowl is an example of a natural La Niña drought event and extreme temperature
event that was exacerbated by human-caused land-use change and poor agricultural practices. These
temperature extremes held records in the Great Plains that have only been surpassed in recent years
(Meehl et al., 2022). Models using only sea-surface temperatures during the 1930s show that the natural
drought would have occurred farther south without human-induced land-use change (Cook et al., 2009).
More recent modeling studies show that the extreme drought of the Dust Bowl caused anomalous
temperatures of +7.9°F (4.4°C) in the Great Plains and of +0.56°F (0.31°C) over the North American
landmass (Meehl et al., 2022). The Meehl et al. (2022) article shows that regional effects like extreme
land-use misuse could have temperature extreme implications not only regionally but for the entire United
States.
A warmer atmosphere increases the maximum amount of water vapor a volume of air can hold at
a specific temperature, which can increase the potential for heavy precipitation even where annual total
precipitation changes little (Eyring et al., 2021; USGCRP, 2023). Warming also increases evaporative
demand (vapor pressure deficit) and shifts snow-to-rain ratios and snowmelt timing, changing soil
moisture and streamflow seasonality. The balance of precipitation versus evapotranspiration 4 determines
drought type and severity.
Atmospheric Moisture
Observations indicate a global increase in surface specific humidity 5 from 1973-2019 (Eyring et
al., 2021; Gulev et al., 2021). This trend is also apparent for the United States (Marvel et al., 2023) and is
consistent with the increased water holding capacity of a warmer atmosphere. Although the specific
humidity is increasing, relative humidity is expected to remain approximately constant as the atmosphere
warms. The increased specific humidity means the atmospheric water vapor content is also increasing
4
Evapotranspiration is the sum of all processes by which water moves from the land surface to the atmosphere via
evaporation (e.g., into the atmosphere from the soil surface or bodies of water on land) and transpiration (water
movement from the soil to the atmosphere via plants).
5
Specific humidity measures the actual mass of water vapor in a unit mass of air. Relative humidity expresses how
close the air is to saturation, as a percentage of the maximum water vapor it can hold at a given temperature; it
therefore varies with temperature, since warmer air can hold more moisture than cooler air.
Prepublication Copy
systematically at a rate of approximately 7% for each degree Celsius of warming experienced. This
increase is both well understood and well documented from global observations (e.g., Santer et al., 2007).
Water vapor accounts for half of the planet’s greenhouse effect and amplifies the effect of GHG-
induced warming. Increased water vapor has a number of important consequences for the hydrological
cycle, in addition to amplifying GHG-induced warming. Storms are producing more intense rains in part
due to this increased water vapor (see following Sections) (Marvel et al., 2023).
Patterns of Precipitation
Over the oceans, multi-decadal salinity analyses show that the spatial patterns of precipitation
have not appreciably changed (e.g., Durack, 2015; Durack and Wijffels, 2010). However, the amplitudes
of the patterns of precipitation and evaporation over oceans have increased, consistent with the “wet-get-
wetter, dry-get-drier” paradigm 6 of change. This is because as temperatures increase, the air holds more
water vapor (up to 7% per degree Celsius of warming), a property known as the Clausius-Clapeyron
relation. Therefore, warmer air increases evaporation over the ocean, holds more water, and increases
heavy rainfall. The implication is that the global hydrological cycle is intensifying, at least over the global
oceans.
Land-based hydrologic trends are complex and not as straightforward as those over the oceans.
Between 1980-2015, water and climate data for the contiguous United States (from National Climate
Assessment Land Data Assimilation System reanalysis), indicate coherent shifts in precipitation across
the country (Jasinski et al., 2019). These shifts reinforce trends noted in previous studies, with mean
precipitation increases of 0.12 to 0.35 inch (3 to 9 millimeters) per year in the upper Great Plains and
Northeast and decreases from −0.04 to −0.35 inch (−1 to −9 millimeters) per year in the West and South.
Patterns of change in terrestrial water storage, revealed by NASA’s GRACE and GRACE-FO satellites,
mirror these precipitation trends. The GRACE data have shown that the areas experiencing drying
globally have increased by twice the size of California annually, creating “mega-drying” regions across
the Northern Hemisphere. While most of the world’s dry regions are becoming drier and wet regions
wetter, the water storage data show that the rate of drying now exceeds the rate of wetting
(Chandanpurkar et al., 2025). The observations of precipitation and water storage are consistent with the
findings from EPA (2009) that “changes are occurring in the amount, intensity, frequency and type of
precipitation.”
Extreme Precipitation
6
The concept that dry regions dry out further, whereas wet regions become wetter as the climate warms has been
proposed as a simplified summary of theoretically expected (e.g., Chou et al., 2009; Held et al., 2006) as well as
observed changes over ocean (Durack et al., 2012), whereas land responses are more complicated (Greve et al.,
2014).
Prepublication Copy
In the United States, the amount of land area experiencing greater than normal annual
precipitation totals has increased since 1895, and the prevalence of extreme single-day precipitation
events has risen substantially since the 1980s (Figure 3.2). These precipitation observations vary both
year-to-year and across different regions of the United States (EPA, 2025d). Regionally, the Northeast has
experienced about a 60% increase in the amount of precipitation falling on the heaviest 1% of days since
1958 (Whitehead et al., 2023), with the Midwest up roughly 45% over similar periods (USGCRP, 2023).
Along the West Coast, landfalling atmospheric rivers 7 have warmed in recent decades (Gonzales et al.,
2019), favoring more rain over snow; multiple reanalysis-based studies detect strengthening characteristics
of atmospheric rivers (Henny and Kim, 2025). Detection-and-attribution analyses using observed records
also identify a human contribution to intensifying 1- and 5-day precipitation extremes (Sun et al., 2021).
Storms
Warmer water temperatures are expected to strengthen tropical cyclones globally (USGCRP,
2023). Although hurricane landfalls in the United States have not increased, hurricane activity in the
North Atlantic has increased since the early 1970s (USGCRP, 2023). A trend has emerged of more rapid
intensification of hurricanes since the early 1980s, as well as a slowdown in the rate of decay of
hurricanes since the 1960s (Kossin et al., 2020; USGCRP, 2023).
Along the North American coast, observations have shown storms slowing down or stalling,
bringing more heavy rainfall, wind damage, storm surge, and coastal flooding (Kossin, 2018; USGCRP,
2023). The destructive power of individual tropical cyclones through flooding is amplified by rising sea
level, which very likely has a substantial contribution at the global scale from anthropogenic climate
change (Knutson et al., 2021). The amount of tropical-cyclone-related rainfall that any given local area
will receive increases as the rain rates at the center of the cyclones increase.
Drought
Drought conditions have also varied over space and time in the United States. Meteorological
droughts (i.e., periods of low precipitation) have increased in the southwestern United States and parts of
the southeastern United States. from 1915-2011. A mixture of positive and negative trends is observed
elsewhere in the contiguous United States (Apurv and Cai, 2021). Analyses that incorporate
evapotranspiration (i.e., evaporation from soils and open water plus plant transpiration) in addition to
temperature to consider soil-moisture drought show a similar pattern with increasing trends in dry area
coverage in the southwest and slight decreasing trends in the rest of the contiguous United States (Su et
al., 2021).
Drought can also be characterized as a sustained imbalance between precipitation and
evaporation. Rising temperatures associated with climate change have accelerated the hydrologic cycle by
increasing evapotranspiration. While greater evapotranspiration places more moisture in the atmosphere
and can enhance precipitation, it also promotes drying over land and reduces soil moisture in many areas.
This pattern is consistent with observations in certain regions: an extended period of drought conditions in
the southwestern United States has been observed from 2012-2023 (EPA, 2025); drought in this region
was also noted in EPA (2009) from 1999-2008.
Since 2009, evidence of these changes has improved through expanded soil-moisture and snow
remote sensing, improved reanalyses for drought process diagnostics, and clearer attribution of heat-
driven aridity to anthropogenic warming in the western U.S. drought signal (USGCRP, 2023; Williams et
al., 2022). Further, human-caused warming has changed the main driver of the soil moisture droughts
over the western United States, from precipitation deficit to heat-driven high evaporative demand, since
2000 (Zhuang et al. 2024).
7
Atmospheric rivers are bands of condensed water vapor in the atmosphere and cause significant levels of
precipitation.
Prepublication Copy
FIGURE 3.2 (Top) Extreme 1-day precipitation events in the contiguous United States, 1910-2023. Percentage of
land area of the contiguous states where much greater than normal portion of total annual precipitation has come
from single-day precipitation events 1910-2023. The bars represent individual years, while the line is a 9-year
weighted average. Data from NOAA NCEI. (Bottom) The frequency and intensity of heavy precipitation events
have increased across much of the United States, particularly the eastern part of the United States, with implications
for flood risk and infrastructure planning. Maps show observed changes in three measures of extreme precipitation:
(a) total precipitation falling on the heaviest 1% of days, (b) daily maximum precipitation in a 5-year period, and (c)
the annual heaviest daily precipitation amount over 1958-2021. Numbers in black circles depict percent changes at
the regional level. Data were not available for the U.S.-Affiliated Pacific Islands and the U.S. Virgin Islands. Figure
credit: NOAA NCEI and CISESS NC.
SOURCE: (Top) EPA, 2025d precipitation indicator: https://www.epa.gov/climate-indicators/climate-change-
indicators-heavy-precipitation. (Bottom) USGCRP, 2023.
Prepublication Copy
The evidence that the ocean has warmed as a result of excess GHGs has grown stronger since
EPA (2009). Because water has a much higher heat capacity than the atmosphere, the ocean is the main
reservoir for heat in the climate system. Ocean warming starts at the surface but is transferred to deeper
layers by ocean circulation. An accurate estimate of ocean heat content is fundamental to understanding
the evolving climate system and fundamental in the estimation of the Earth energy imbalance (Section
2.3). A robust increase in global sea surface temperature has been observed since 1900 (Garcia-Soto et al.,
2021) and increases in summer upper-ocean stratification are apparent from 1970-2018 (Sallée et al.,
2021). Increases in the heat content of the upper approximately 6,560 feet (2,000 meters) of the ocean are
also evident since 1960 throughout the global oceans (Garcia-Soto et al., 2021), consistent with the
Earth’s energy imbalance.
The advent of the Argo float network in 2004 greatly improved the spatial and temporal coverage
of in situ measurements of temperatures in the upper layer of the ocean, which previously had been
measured by electronic instruments lowered from ships. Currently, more than 3,900 Argo floats provide
about 140,000 temperature (and salinity) profiles per year from the sea surface to about 6,560 feet (2,000
meters) depth at places across the globe (NASEM, 2017). The Deep Argo program, which began in 2014
and expanded in 2016, advanced sampling of temperatures down to about19,685 feet (6,000 meters) depth
and enabled estimation of ocean heat gain over the full water column.
FIGURE 3.3 Annual ocean heat content compared to the 1993 average from 1993-2019, based on multiple data
sets: surface to depths of 2,300 feet (700 meters) in shades of red, orange, and yellow; from 2,300-6,650 feet (700-
2,000 meters) in shades of green and blue; and below 6,650 feet (2,000 meters) as a gray wedge. Graph by NOAA
Climate.gov, adapted from Figure 3.6 in State of the Climate in 2019.
SOURCE: NOAA, 2025.
Prepublication Copy
Heat absorbed by surface ocean waters is transported laterally and vertically through the layers
and basins of the ocean via mixing and currents. Regionally, subsurface ocean temperature can also vary
substantially with climate patterns such as El Niño, the Pacific Decadal Oscillation, the North Atlantic
Oscillation, and large variations in wind stress over the ocean. On a regional basis, closure of the heat
budget requires observations of ocean heat content, air-sea heat exchange, heat transport by ocean
currents, and mixing.
Ocean warming has contributed to increases in rainfall intensity, rising sea levels due to thermal
expansion, the destruction of coral reefs, declining ocean oxygen levels, and declines in ice sheets,
glaciers, and ice caps in the polar regions (Cheng et al., 2019; Hamlington et al., 2022). This warming is
also one of many factors that has increased the number of low-oxygen dead zones in many places around
the United States (USGCRP, 2023).
Marine heat waves are periods of anomalously high regional surface ocean temperatures. These
events were not included in the evidence in EPA (2009), but have become common in recent decades.
These heat waves have considerable and detrimental impacts on marine ecosystems and the services that
they provide (e.g., Frölicher and Laufkötter, 2020; Smale et al., 2019; Smith et al., 2024). In 2024, 91%
of the global ocean was affected by at least one marine heat wave, while 26% experienced at least one
cold spell (Blunden and Reagan, 2025). Laufkötter et al. (2020) show that the frequency of these events
has increased more than 20-fold since preindustrial times, when it is estimated that they typically occurred
with a frequency of once in hundreds to thousands of years. From 1982 to 2023, the annual cumulative
intensity of marine heat waves increased across almost all waters in the U.S. exclusive economic zone,
except for a nearshore region from Georgia to North Carolina (see Figure 3.4; EPA, 2025e). The largest
changes are present in Alaskan coasts and waters off the northeastern United States.
Ocean Chemistry
The continued observations of ocean chemistry since 2009 affirm the conclusions in EPA (2009)
about uptake of excess CO2 from the atmosphere. The ocean has taken up about 30% of the CO2 emitted
to the atmosphere over the past century (Gruber et al., 2019). Uptake of CO2 in the ocean leads to a series
of chemical reactions that lower the pH, increase concentrations of dissolved organic carbon, and increase
the solubility of calcium carbonate. Calcium carbonate is an important component of the shells and
skeletons of many marine organisms. EPA (2009) noted that ocean pH had decreased from 8.2 in 1750 to
8.1 today, following a trend that has continued since 2009. Because the pH scale is logarithmic, a 0.1-unit
decrease represents about a 30% increase in the hydrogen ion concentration in ocean water, which makes
it less alkaline. The Intergovernmental Panel on Climate Change (IPCC) has assessed that is virtually
certain that human-caused CO2 emissions are the main driver of the current global decline in pH of the
surface open ocean; a pH decline in the ocean interior over the past 2-3 decades has also been observed in
all ocean basins (with high confidence; IPCC, 2021). The decline in pH in U.S. offshore waters tracks
with the global average trends, but changes in U.S. coastal waters vary due to upwelling conditions and
nutrient and freshwater inputs that also lower the pH of ocean water (USGCRP, 2023).
Changes in ocean pH are monitored through in situ measurements of pH and partial pressure of
CO2 (a measure of the quantity of CO2 dissolved in seawater). The number of moored and shipboard
sensors for pH and CO2 have increased greatly over the past decade and now provide more accurate
monitoring information than in 2009. The rise in atmospheric CO2 concentration concurrent with in situ
measurements of pH and the partial pressure of CO2 in ocean water clearly illustrate a cause-and-effect
relationship between these variables over the long-term. Ma et al. (2023) used in situ and satellite
observations to examine the trend in ocean pH from 1982-2021, confirming that the declining pH across
the global ocean is attributable to the increase in the partial pressure of CO2 from human-caused increases
in atmospheric CO2. The IPCC estimates in the Special Report on the Ocean and Cryosphere in a
Prepublication Copy
Changing Climate (IPCC, 2019) that the rate of ocean surface pH decline is 0.017-0.027 pH units per
decade across a range of time series that are longer than 15 years. Bates and Johnson (2020) found that
seawater CO2-carbonate chemistry conditions today clearly exceed seasonal changes observed in the
1980s.
FIGURE 3.4 Change in annual cumulative intensity of marine heat waves in the United States, 1982-2023.
Cumulative intensity is measured in degree days—marine heat wave intensity multiplied by duration. Areas with
increases are shown in red, with darker colors indicating greater change. The map shows total change, which is the
annual rate of change (trend slope) multiplied by the number of years analyzed. Data from NOAA NCEI.
SOURCE: U.S. Environmental Protection Agency. Climate Change Indicators in the United States. Marine Heat
Waves. Accessed September 2025. https://www.epa.gov/climate-indicators/climate-change-indicators-marine-heat-
waves.
The study of ocean pH decline and its effects on marine organisms has expanded dramatically
over the last two decades (Browman, 2016), and research demonstrates varied responses among various
communities and species, with calcifying species (including corals) generally exhibiting more sensitivity
to higher CO2 (Doney et al., 2020; Kroecker et al., 2013). The extent to which the effects of ocean pH
decline on marine biota will impact human welfare is an area of active research. Some areas of focus
include potential impacts on economics of commercial fisheries and tourism, cultural values, and role in
coastal protection for corals (Doney et al., 2020). See Chapter 6 for further discussion of impacts.
Prepublication Copy
EPA (2009) documented a number of changes in physical and biological systems, including in the
cryosphere, hydrosphere, and biosphere. The report highlighted the finding from IPCC (2007) that
“anthropogenic warming has had a discernible influence on many physical and biological systems,” but
also noted that other factors, such as land-use change or natural decade-scale climate variations (such as
the Pacific Decadal Oscillation) were likely to play a role.
Since EPA (2009), data records have lengthened, data coverage has improved for some variables
(e.g., mountain glaciers), new data sources have become available (e.g., ICE-Sat2 and GRACE satellites),
and improved methodologies have been devised for assessing change. The changes in physical and
biological systems documented in 2009 have generally continued and in some cases become more clearly
attributable to a human influence (IPCC, 2019, 2021).
Cryosphere
The trend in annual mean Arctic sea ice extent remains similar to that documented in EPA
(2009), with a loss per decade of ~500,000 square-kilometers (about 4.5% relative to the 1981-2010
average) for the period of 1979-2023 (Fetterer et al., 2025). Sea ice has decreased in all months relative to
the historical average, with the largest reductions in September equaling about 12.2% per decade for
1979-2023 relative to the 1981-2010 period. The melt season has also lengthened from 1979-2023, with
both earlier melt onset and later freeze-up (EPA, 2025b). A pause in September sea ice loss has occurred
in the last two decades. This pattern is consistent with internal climate variability (England et al., 2025)
and was anticipated in work showing that decadal pauses in ice loss are possible when anthropogenic ice
loss is counteracted by internal variability (Kay et al., 2011).
EPA (2009) noted that for the 1979-2008 period, Antarctic sea ice exhibited no significant
change. However, since that time, Antarctic sea ice has undergone a significant loss. A small but
significant increase in ice extent occurred from 1979-2014. This was followed by a dramatic ice loss in
austral spring of 2016. Ice extent has remained remarkably low during the last decade with losses
comparable to the 46-year record of Arctic sea ice decline (Abram et al., 2025). This has led to the
suggestion that a regime shift may have occurred in the Antarctic sea ice system (Hobbs et al., 2024;
Purich and Doddridge, 2023).
The other cryospheric changes documented in EPA (2009) have generally continued. For the
1961-2016 period, glacier mass has been lost globally (Zemp et al., 2019), with glaciers in Alaska losing
3000 Gigatons (equivalent to 0.31 inches, or 8 millimeters, of sea level rise) and in western Canada and
the western United States losing 428 Gigatons (equivalent to about 0.04 inches, or 1 millimeter, of sea
level rise) of mass. Permafrost continues to warm and thaw, and lake ice cover has declined (IPCC, 2019;
Vonk et al., 2015). Northern Hemisphere spring snow cover has continued to decline, with a loss since
1922 of approximately 0.3 million square-kilometers per decade (IPCC, 2021).
Sea Level
As discussed in the previous sections, Earth’s energy imbalance has led to warming of the
surface, and 90% of the excess heat has been absorbed by the oceans. This has led to thermal expansion
of the oceans which, together with land ice mass loss, contribute to sea level rise. Global mean sea level
has risen about 7 inches (approximately 18 centimeters) since 1900, up from 6.7 inches reported in EPA
(2009).
Over the past three decades, satellites have provided continuous, accurate measurements of sea
level on near global scales. EPA (2009) reported the global average rate of sea level rise from 1993-2003
as measured by satellite altimetry to be about 0.12 inch (3.1 millimeters) per year; recent studies show the
rate increased from about 0.08 inch (2.1 millimeters) per year in 1993 (the first year in previous averaged
period) to about 0.18 inch (4.5 millimeters) per year in 2023 (Hamlington et al., 2024). This acceleration
Prepublication Copy
of the rate of sea level rise has been evident in both tide gauge and satellite altimetry records (Eyring et
al., 2021; Sweet et al., 2022).
Regional relative sea level (see Figure 3.5) rose on average by approximately 11 inches (28
centimeters) in the last century along the continental United States, with about half of this amount (5-6
inches, about 13-15 centimeters) in the last 30 years (USGCRP, 2023). The greatest rise in regional
relative sea level during that time (9 inches, about 23 centimeters) was along the U.S. western Gulf Coast,
largely due to land subsidence from groundwater and fossil fuel extraction. Comparatively, regional
relative sea level rise along the northeast and southeast Atlantic and eastern Gulf coasts was
approximately 6 inches (about 15 centimeters). Along the Pacific Coast, natural modes of variability (the
El Niño-Southern Oscillation and Pacific Decadal Oscillation) will continue to drive decadal variability in
the rate of sea level rise.
FIGURE 3.5 Observed sea level trends in the United States, 1993-2020. Data from tidal gauges and satellites show
that on average the United States sea level rise trends are higher than the global average. The highest rates of sea
level rise are shown in red and occur on the Atlantic Coast and Gulf region while low rates of sea level rise are
shown in blue and occur along parts of the Pacific and Alaskan Coast.
SOURCE: Marvel et al., 2023.
Risks from rising sea levels include increased coastal flooding and increased vulnerability to
coastal storms. Changes in average sea level have doubled the frequency of high tide flooding in the
continental United States over the past few decades (USGCRP, 2023). In some cities, higher rates of local
sea level rise have increased flood frequency. For example, using data from 1998-2013, Wdowinski et al.
(2016) showed that significant changes in Miami Beach, Florida, flood frequency occurred after 2006,
with four times as many disruptive high tide flooding events compared to the 1998-2005 period.
The longer observational records have increased confidence in estimates of human-caused sea
level rise, and acceleration in the rate of increase. Furthermore, with benefit of a longer record, patterns
and influences of natural variability of the ocean on interannual to decadal timescales can be more readily
Prepublication Copy
identified. This strengthens confidence since EPA (2009) to support the conclusion that “global sea level
gradually rose in the 20th century and is currently rising at an increased rate.”
Biosphere
Ground-Level Ozone
As EPA (2009) described, climate change affects ground-level ozone by modifying precursor
emissions (other compounds that produce ozone through chemical reaction in the atmosphere),
atmospheric chemistry, transport, and removal. USGCRP (2023) expanded on climate-sensitive factors
driving increases and decreases in ozone concentrations, which include temperatures, heat waves,
wildfires, drought, and biogenic emissions. Climate change can also reduce ozone pollution through
increasing humidity. Climate change-induced changes in precipitation do not affect ozone, and effects of
climate change-induced changes in regional transport and stagnation are unknown (USGCRP, 2023).
New studies further corroborate the effects of climate change on ground-level ozone reported in
EPA (2009) (USGCRP, 2016, 2023). Historical climate change has increased peak season ozone
concentrations over North America (Turnock et al., 2025). Ozone increases driven by climate change may
put some areas of the United States into nonattainment with the ozone National Ambient Air Quality
Standard (Chang et al., 2025; East et al., 2024). Wildfires release ozone precursor gases and therefore can
worsen ozone concentrations (Cooper et al., 2024), including far downwind of the fire; this is
demonstrated by, for example, fires in Canada increasing ozone in Midwest United States (Cooper et al.,
2024) and fires in California, Washington, and Arizona increasing ozone in northern Colorado by 8 parts
per billion in July 2021 (Langford et al., 2023). Studies find increasing co-occurrence of ground-level
ozone with fine particulate matter, wildfire smoke, and heat extremes (Kalashnikov et al., 2022). Ozone,
heat, and particulate matter can act synergistically on biological systems, leading to worse health
outcomes compared with ozone exposure alone (see Chapter 5; Anenberg et al., 2020; Fann et al., 2021;
He et al., 2025; Remigio et al., 2021).
In addition to the impacts of climate change on ground-level ozone, CH4 is both a potent GHG
and a precursor for ground-level ozone. While local, daily variations in ozone concentrations are largely
driven by reaction of non-CH4 volatile organic compounds with nitrogen oxide emissions in the presence
of sunlight, decades of research show that CH4 also influences long-term average ozone concentrations
globally and in the United States (Fiore et al., 2002, 2008; McDuffie et al., 2023; Shindell et al., 2024;
West and Fiore, 2005; West et al., 2006; Zhang et al., 2016). One study estimated that global CH4
increases contributed 15% of observed trends in daily maximum 8-hour average ozone over the western
United States from 1980-2014 (Lin et al., 2017).
Prepublication Copy
3.6 WILDFIRES
The evidence supporting the EPA (2009) discussion of impacts of climate on wildfires has
strengthened greatly since 2009, as the occurrence of wildfires in the western United States has increased
(Abatzoglou and Williams, 2016; Duffy et al., 2018). While records show a deficit in widespread fire
relative to pre-1880 fire regimes, this is largely a result of fire suppression (Parks et al., 2025). Despite
this deficit, wildfire intensity has been amplified by climate change (Jones et al., 2022; Parks and
Abatzoglou, 2020), even under human fire suppression.
Increasing wildfire severity and area burned are linked to climate change (USGCRP, 2023).
Warmer springs, prolonged summer dry periods, and progressively drier soils and fuel sources increase
the likelihood for wildfire ignition and spread (Ostoja et al., 2023). Earlier snowmelt and diminished
snowpack reduce water availability during peak summer heat, further lowering fuel moisture and enabling
hotter, more intense burning. These dynamics are expected to persist as droughts become more frequent
and longer in duration in some regions within the United States. Since 2009, these same drivers—earlier
spring onset, longer summer dryness, and cumulative drought stress—have continued to lengthen the fire
window and raise the probability of large, fast-spreading events, especially in the western United States.
Observations have been consistent with these mechanisms. In the West, both total burned area
and the area burned at high severity have increased alongside warmer, drier fire seasons and higher vapor-
pressure deficit (a measure of fuel aridity) (Abatzoglou and Williams, 2016). Synthesizing satellite burn-
severity maps with incident records indicates roughly an eightfold rise in annual area burned and in high-
severity burned area in western forests since the mid-1980s (EPA, 2025g). National indicators show
greater acreage burned and longer fire seasons in the West relative to the East.
With increased wildfires, substantial amounts of particulate matter are produced (Law et al.,
2025). Exposure to fine particulate matter is a known cause of mortality and cardiovascular disease and is
linked to onset and worsening of respiratory conditions (see Section 5.3).
Wildfire Feedbacks
Changes in the water cycle are making forests in the western United States more susceptible to
drought and wildfire. Moreover, increases in atmospheric nitrogen deposition and ground-level ozone
shift the processing of water, carbon, and nitrogen in forest ecosystems, resulting in a cascade of
synergetic effects that make trees more prone to disease, pest invasion, drought, and ultimately wildfire.
These air pollutants increase leaf turnover and litter mass and decrease the decomposition of litter
(Gilliam et al., 2019). As a result, mixed conifer forests of southern California that are impacted by
nitrogen and ozone pollution develop deep litter layers. Elevated ozone decreases plant control of water
loss, increasing transpiration, which when coupled with loss of root mass, increases the susceptibility of
trees to drought stress and makes them more vulnerable to attack by bark beetles, leading to significantly
higher tree mortality (Jones et al., 2004b).
The enhanced fuel load from tree decline and litter accumulation, coupled with historical fire
suppression, increases in housing developments at the wildland-urban interface, and climate change, have
resulted in catastrophic fires in northern and southern California in recent years, exacerbating air quality
and health impacts. Such feedbacks are expected to continue to influence wildfires in the future, which
would accelerate a deterioration in air quality and its associated impacts.
Wildfires also release large amounts of CO2, CH4, and other GHGs, as well as black carbon
particles into the atmosphere, which contribute to climate warming, leading to a positive feedback loop
that could further increase wildfire risk (NASEM, 2024d). Black carbon, a potent short-lived climate
forcer present in wildfire smoke, accelerates glacier and snow melt and amplifies atmospheric warming.
Elevated emissions from recent wildfires have been measured at levels equivalent to the annual fossil fuel
output of major industrialized nations (Byrne et al., 2024). Jones et al. (2024b) found that global CO2
emissions from forest fires have surged by 60% since 2001 largely due to increasingly intense and wide-
ranging wildfires.
Prepublication Copy
8
In this context, “whiplash” refers to rapid swings from one extreme or specified climate state to another (e.g., from
drought to flood or from frozen to unfrozen winter weather conditions).
Prepublication Copy
4
Impacts of Greenhouse Gas Emissions on Future Climate
EPA (2009) provided projections of future changes in the climate system associated with human-
caused greenhouse gas (GHG) emissions. Many of these projected changes have been observed since
2009, as described in Chapter 3, including increasing surface temperatures, higher sea levels, and regional
variability across the United States in other physical and biological systems.
Continued emissions of greenhouse gases from human activities will lead to more climate
changes in the United States, with the severity of expected change increasing with every ton of
greenhouse gases emitted. Despite successful efforts in many parts of the world to reduce emissions,
total global GHG emissions have continued to increase, and additional warming is certain.
Models have proven skillful and are effective at simulating a fingerprint of human influence
on the changing climate that is now observed. Climate models, which simulate the Earth system, have
been used since the 1960s to examine the role of different climate forcings in driving climate variability.
Models have simulated certain “fingerprints” of the climate response to human-caused GHG emissions
that have since been observed, including the vertical structure of temperature changes and enhanced
warming over land relative to oceans.
All climate models—regardless of assumptions about future emissions scenarios or
estimates of climate sensitivity—consistently project continued warming in response to future
atmospheric greenhouse gas increases. Projections of future change draw primarily on physically-based
climate models, which have advanced in spatial resolution, process representation, and evaluation since
2009, improving confidence in understanding of the implications of future emissions. Applying
fundamental physics of the Earth system leads to the same conclusion about future warming as projected
by climate models.
Continued changes in the climate increase the likelihood of passing thresholds in Earth
systems that could trigger tipping points or other high impact climate surprises. These surprises are
difficult to predict, can occur abruptly, and, in some cases, would be irreversible.
Climate models are numerical simulations of the Earth system, including the atmosphere, ocean,
land, freshwater systems, and sea ice, and the coupling among these components. These models are based
on the underlying physics that govern these systems and evolved from numerical weather prediction
models, first developed in the 1960s (e.g., Manabe and Wetherald, 1967). Aided by improvements in their
process 1 representation, as well as improvements in supercomputers, these models have increased in
complexity. These models are now often referred to as “Earth System Models” and can incorporate many
additional components and processes, including terrestrial and marine ecosystems, atmospheric
chemistry, land ice, and glacial dynamics.
Since the 1960s, climate models have been used to examine climate variations and the role of
different climate forcers, such as GHGs and volcanic emissions, in driving that variability. They provide
projections of future climate conditions subject to scenarios of future emissions of GHGs and aerosols, as
well as land cover changes. They are also an important tool for attribution in that they allow for controlled
1
In climate and Earth system models, a “process” refers to a physical, chemical, or biological phenomenon—such
as cloud formation, ocean circulation, or carbon uptake by plants—that influences the climate system and is
described mathematically in the model.
Prepublication Copy 36
experiments, for example by isolating the climate response to increasing GHGs versus other climate
drivers (e.g., Gillett et al., 2016). Notably, these models have improved at simulating changes in the
climate that have already been observed.
However, the climate system is complex, and models are imperfect tools. Climate projections
have uncertainty due to internal variability in the climate system, uncertainty in future emissions of GHGs
and aerosols, and structural uncertainty in the models themselves. Comparisons across models, which
include different details of process representation and numerical implementation, are useful for
quantifying structural model uncertainty (e.g., Hawkins and Sutton, 2009). A common metric used for
this comparison is the equilibrium climate sensitivity (see Jeevanjee et al., 2025, for a recent review),
which is the equilibrated global mean surface warming for a doubling of CO2. A subset of climate models
have equilibrium climate sensitivity that is higher than the likely range as assessed by the
Intergovernmental Panel on Climate Change (e.g., Zelinka et al., 2020). Understanding why these models
have a higher equilibrium climate sensitivity is an area of active research. That said, while this metric is
useful for understanding climate response and model uncertainty, it has somewhat limited relevance for
projected change in the near term because the climate system is not equilibrated and factors such as
changing aerosol emissions play a role (Jeevanjee et al., 2025).
Since EPA (2009), updates to climate models have occurred with the newest models available
under the Coupled Model Intercomparison Project 6 (Eyring et al., 2016). As a group, these models have
improvements to process representation and include additional capabilities. For example, advances have
been made since 2009 in the ability for model ensembles to quantify the influence of internal climate
variability on projections (Kay et al., 2015) and in model resolution. These improvements have enabled
the use of models for new applications, such as multi-year prediction of flood frequency (Zhang et al.,
2025) and extending the time horizon of tropical cyclones forecast to seasons (Murakami et al., 2025).
The availability of longer observational time series since EPA (2009) has also allowed for
improved validation of model-simulated trends in the historical record and improved understanding of
model successes and challenges that are still present (Simpson et al., 2025). The models have allowed the
detection of a “fingerprint” of human influence (see Section 2.4) across many observed changes in the
Earth system (Eyring et al., 2021), including the vertical (Santer et al., 1996) and regional (Hegerl et al.,
1996) structure of temperature changes, seasonal cycle changes for tropospheric (Santer et al., 2022) and
sea surface (Shi et al., 2024) temperatures, and daily precipitation variability (Ham et al., 2023), among
others.
While models are not perfect, they are useful and skillful tools for attribution of anthropogenic
signals in the changing climate and understanding of future climate changes in response to GHG
emissions. All climate models consistently project continued warming in response to future GHG
increases, regardless of climate sensitivity level or future emission scenario. Notably, they are just one
line of evidence of human influence on historical climate change. When combined with observational
evidence, paleoclimate information, and theoretical understanding, it is unequivocal that many climate
changes underway can be attributed in large part to rising GHG emissions from human activity. This
evidence also indicates that every additional quantity of emissions will strengthen, and in some cases
accelerate, those changes for the future.
The long lifetime of energy, transportation, industrial, and other built infrastructure creates
challenges in rapidly reducing CO2 emissions (NASEM, 2021, 2024). Hence, despite mitigation efforts in
many parts of the world and the achievement of downward CO2 emission trends in many advanced
economies, total global emissions have continued to grow and are expected to remain near current levels
over the coming decade.
Given the persistence of CO2 and other long-lived GHGs in the atmosphere, past emissions have
increased atmospheric concentrations of these gases, which will sustain Earth’s energy imbalance
(Section 2.3). Warming will continue until net CO2 plus N2O emissions reach (and remain at) ~zero and
Prepublication Copy
emissions of CH4 are constant or decreasing. As long as global emissions of CO2 stay above zero,
concentrations and radiative forcing will continue to increase and global temperature will increase
roughly in proportion to cumulative CO2 (i.e., each additional ton emitted adds an increment more to
temperature increase) with small contributions from other long-lived gases including N2O and F-gases. As
global emissions of GHGs are spread across all nations, a collective effort at reducing emissions is
required to limit future warming.
Analysis of policies in place in 2023 showed that CO2e emissions would stay roughly constant
over 2025-2035 (UNEP, 2024), driving continued warming that would lead to a projected peak global
mean warming in 2100 of about 4.9°F (2.7°C) (4.1 to 5.4°F, or 2.3 to 3°C range) (Hausfather, 2025).
Were all countries to fully implement their 2023 policies and their unconditional pledges to the United
Nations Framework Convention on Climate Change, global CO2e emissions would drop by ~5%, leading
to a projected peak 21st-century warming of about 4.4°F (2.4°C), or only about 0.5°F (0.3°C) less than
under current policies (UNEP, 2024). The likelihood of exceeding global mean 3.6°F (2°C) warming
relative to preindustrial temperatures under these two cases is estimated at 97% with current policies
continuing and 94% for unconditional pledges continuing (UNEP, 2024). Warming beyond 3.6°F (2°C) is
expected to have many negative consequences for human health and welfare in the United States (see
Chapters 5 and 6).
Emission scenarios used in simulations of projected climate (O’Neill et al., 2016; Riahi et al.,
2017) have been updated since EPA (2009). These scenarios encompass a range of possible futures,
including high emissions scenarios, which assume significant “regional rivalry,” and “sustainability”
scenarios, which assume deep and sustained reductions in emissions, with net negative CO2 emissions by
2100. Using this range of scenarios and large numbers of model simulations, projected impacts of future
warming on the United States can be assessed probabilistically. Risks of future impacts for some
quantities—including heat, sea level rise, and some extreme events—can be assessed with relatively high
confidence, while risks for other expected impacts, including regional droughts and hurricane intensities,
continue to have large uncertainties in their quantification. Climate-related damages increase with every
quantity of GHGs emitted and the damages per ton of emissions also rise as the Earth continues to warm
owing to non-linearities in impacts (Cissé et al., 2022; EPA, 2023). In simulations with sustained GHG
emission reductions, the increase in atmospheric CO2 concentrations slows after 5-10 years and global
warming slows after several decades (Lee et al., 2021). This is consistent with studies that examine the
climate response to a complete cessation of CO2 emissions, which indicate that temperatures would
stabilize or even decrease over time (Jones et al., 2019; Matthews and Weaver, 2010).
With each increment of continued GHG emissions and warming, surface and near-surface air
temperatures (and thus heat exposure to humans and e.g., crops, animals, and ecosystems) increase;
extreme heat becomes more frequent and extreme precipitation events increase across some regions,
while aridification and drought persist in others—patterns that often scale approximately linearly with
global temperature, though not uniformly across all metrics or places (USGCRP, 2023). The oceans
continue taking up heat and CO2, driving higher ocean heat content, rising sea levels from thermal
expansion and land-ice loss, and decline in ocean pH; these changes persist for decades to centuries even
if temperatures stabilize (Lee et al., 2021). These findings are independent of the equilibrium climate
sensitivity level in any specific model.
In response to increasing GHGs, many climate changes have been observed and are anticipated
for the future. In many cases, these changes exhibit a roughly linear relationship to changed forcing, as
with global temperature, discussed above. However, the climate system can also exhibit an abrupt
response to climate warming when certain thresholds (sometimes referred to as “tipping points”) are
passed. Paleoclimate data indicate that abrupt shifts have occurred in the past (e.g., Capron et al., 2021).
With continued and accelerating climate warming, the likelihood for surpassing thresholds grows with the
potential for rapid and dramatic disruption to human systems (NRC, 2013).
Prepublication Copy
EPA (2009) states, “Climate warming may increase the possibility of large, abrupt regional or
global climatic events.” This was and remains accurate, supported by more evidence on additional
possible “tipping elements” that could undergo abrupt change. EPA (2009) discussed several specific
elements where abrupt changes were possible, including megadroughts, disintegration of the Greenland
Ice Sheet, collapse of the West Antarctic Ice Sheet, catastrophic release of CH4 from sea floor CH4-
hydrates and/or permafrost soils, and slowing down of the Atlantic Meridional Overturning Circulation
(AMOC)—which is a major component of the Atlantic ocean circulation. For many of these elements,
abrupt change in the 21st century was considered to be low probability but high impact. There remains
uncertainty in whether or when tipping points might be reached in these elements. For example, studies
on AMOC simulate a range of responses to changing buoyancy (heat and freshwater) forcing (e.g.,
Jackson et al., 2023) and observationally-based early warning systems may suggest a higher likelihood of
collapse than seen within climate models (e.g., Ditlevsen and Ditlevsen, 2023). Paleoclimate evidence
strongly suggests instances of AMOC collapse during the Younger Dryas, Heinrich events, and
potentially during Dansgaard–Oeschger 2 events in the last glacial period (e.g., Lynch-Stieglitz, 2017).
Since 2009, evidence has emerged for some abrupt changes underway. For example, numerous
rapid changes in the Antarctic environment have occurred (Abram et al., 2025), including rapid
reductions in sea ice (Purich and Doddridge, 2023), regime shifts in biological systems (e.g., Fretwell et
al., 2023), and increasing ice sheet mass loss (Rignot et al., 2019), with consequences for sea level rise.
Research has also highlighted potential additional “tipping elements,” including rapid changes in
numerous terrestrial and marine ecosystems, expansion of oxygen minimum zones, the potential collapse
of the Antarctic Overturning circulation (Abram et al., 2025), and loss of alpine glaciers, among others
(e.g., NRC, 2013). Assessments have quantified an increasing likelihood of passing multiple climate
tipping points with increasing warming (e.g., Armstrong McKay et al., 2022), many of which could be
irreversible. Work has also highlighted that tipping elements can interact and often do so in destabilizing
ways, thus setting up the possibility of “tipping cascades” (Wunderling et al., 2024).
2
Dansgaard–Oeschger (D-O) events are periods of rapid warming (over a few decades) followed by a slow cooling
(over a few hundred years). During the Last Glacial Period there were 25 recorded D-O events that occurred every
few thousand years.
Prepublication Copy
5
Impacts on Human Health
Human-caused emissions of greenhouse gases and resulting climate change harm the health
of people in the United States. Evidence since 2009 supports and strengthens EPA (2009) conclusions
and has deepened the understanding of how these risks affect health. Climate-related illnesses and deaths
are increasing in both severity and geographic range across the United States.
Climate change intensifies risks to human health from exposures to extreme heat, ground-
level ozone, wildfire smoke and other airborne particulate matter, extreme weather events, and
airborne allergens, affecting incidence of cardiovascular, respiratory, and other diseases. Much
evidence is now available on how heat affects morbidity and mortality, not only by directly causing heat
exhaustion and heat stroke, but also by worsening effects on cardiovascular, respiratory, kidney, mental
health, and other disorders. New evidence has deepened understanding of how climate-sensitive drivers
increase ozone pollution, how long-term ozone exposure leads to health effects beyond those of short-
term exposure, and how health outcomes are amplified by co-occurrence of ozone exposure with heat and
particulate matter exposure.
Climate change has increased exposure to wildfire smoke and dust, which has been linked
to adverse health effects. Since 2009, wildfire smoke exposure has increased, particularly in the U.S.
West, and new evidence has linked wildfire smoke exposure to a wide range of adverse human health
outcomes, including respiratory disease and premature death. New evidence has also shown that climate
change has increased airborne soil dust and associated health effects, particularly for areas that are
warmer and drier, such as the U.S. Southwest.
The increasing severity of some extreme weather events, such as wildfires and heavy
precipitation events, has contributed to injury, illness, and death in affected communities. Although
non-climate factors, including adaptation, can mitigate the negative effects of climate change on health,
extreme events can overwhelm the ability to respond. Extreme events can impact human systems,
including health care and food systems, power systems, and other critical infrastructure, adding to the
risks faced by individuals and communities.
Health impacts related to climate-sensitive infectious diseases—such as those carried by
insects and in contaminated water—have increased. An increase in the geographic distribution of tick-
borne diseases, anticipated in the 2009 report, has been confirmed and attributed to climate warming.
Dengue, a mosquito-borne viral disease, has increased in activity and geographic range since the 2009
report, now appearing in non-travelers in Texas, Florida, Arizona, and Hawaii. Climate change is
expanding the area of endemicity of some fungal diseases. Heavy precipitation, drought, and warming
temperatures are linked to a rise in waterborne disease outbreaks.
New evidence is developing about additional health impacts of climate change. Newer areas
of evidence include potential impacts on mental health, nutrition, immune health, antimicrobial resistance,
kidney disease, and negative pregnancy-related outcomes. In addition, research has grown showing that
combined exposure to multiple climate-sensitive risk factors, either simultaneously or cumulatively over
time, worsens health outcomes.
Groups such as older adults, people with preexisting health conditions or multiple chronic
diseases, and outdoor workers are disproportionately susceptible to climate associated health
effects. New findings also point to elevated risks for pregnant people and children. Even as non-climate
factors, including adaptation measures, can help people cope with harmful impacts of climate change,
they cannot remove the risk of harm.
Prepublication Copy 40
Weather and climate interact with many factors to shape the effects of climate-sensitive health
outcomes in any given place and time (Ebi et al., 2020). The climate changes discussed in preceding
chapters affect human health directly through events such as heat waves or wildfires, as well as indirectly,
through pathways like air and water quality and nutrition, with these impacts further shaped by broader
environmental, social, and public health conditions (WHO, 2018). The climate and health field has
expanded considerably since EPA (2009) was published, with substantial new data and research
strengthening the evidence base, deepening understanding of how climate change affects health, and
clarifying the pathways through which these impacts occur. This chapter explores health effects from
exposure to extreme temperatures and events such as wildfires and hurricanes, the influence of climate on
infectious and noncommunicable diseases, and health implications of changes to air quality. Examples
illustrate areas where knowledge has grown or evolved, including areas of health research not discussed
in EPA (2009).
EPA (2009) concluded that negative health effects from climate change are experienced
disproportionately by some populations. Subsequent assessments of the impacts of human-caused
greenhouse gas (GHG) emissions and resulting climate changes have also concluded that these changes
are harming physical and mental health (IPCC, 2022a; USGCRP, 2023; see discussion of GHG emissions
and climate effects in Chapters 2 and 3). The evolution of understanding of health risks from climate
change since 2009 is summarized in Table 5.1. Evidence has continued to accumulate about multiple
health effects; none of the areas identified in EPA (2009) showed weakened evidence of health effects.
Although this report does not consider the impacts of adaptation or mitigation measures in reducing
climate-associated health risks in detail, Box 5.1 highlights a few potential ways such adaptations may
influence risk as well as some of the potential limitations of such measures.
EPA (2009) concluded that “[s]evere heat waves are projected to intensify in magnitude and
duration over the portions of the United States where these events already occur, with potential increases
in mortality and morbidity, especially among the elderly, young, and frail.” Observations continue to
show a warming trend in regions of the United States, particularly in Alaska, the West, and the Northeast,
and hot extremes increasing with cold extremes decreasing (see Section 3.2). Studies and assessments of
human health consequences continue to support the EPA (2009) conclusion that changes in average
temperatures and increased exposure to temperature extremes contribute to adverse health outcomes in
many places in the United States.
Studies on ambient temperature and health have identified U-, V- or J-shaped patterns in which
extremes of both hot and cold are associated with adverse effects, recognizing that duration, humidity
level, extent of divergence from a location’s usual temperature, and other parameters influence these
outcomes and that some groups are more susceptible to the effects of temperature than others (Burkart et
al., 2021; Ye et al., 2012). For example, a 2023 study spanning 2000-2016 and involving 61.6 million
Medicare beneficiaries aged 65 and older found that cardiovascular hospitalizations were higher in areas
with hotter summers or colder winters (Klompmaker et al., 2023). USGCRP (2016) identified as a key
finding that “days that are hotter than usual in the summer or colder than usual in the winter are both
associated with increased illness and death [Very High Confidence]. Mortality effects are observed even
for small differences from seasonal average temperatures [High Confidence].”
Recent evaluations and assessments have highlighted the wide-ranging health consequences of
exposure to increased temperature and heat (IPCC, 2022a; Lancet Countdown 2024; USGCRP, 2023).
Heat contributes to excess illness and death in the United States and globally, with an estimate of net
increases in all-cause mortality risk associated with increased average annual temperatures from 0.1% to
Prepublication Copy
1.1% per 1.8°F (1°C) (Cromar et al., 2022). According to data from the National Weather Service, heat is
associated with more weather-related deaths than any other extreme weather event. 1
Much evidence is now available on how heat affects morbidity and mortality, not only by directly
causing heat exhaustion and heat stroke, but also by worsening effects on cardiovascular, respiratory,
kidney, mental health, and other disorders (see Section 5.5 discussing climate-sensitive diseases). Heat
waves are linked with higher rates of emergency department visits, hospitalizations, or deaths for such
conditions (Khatana et al., 2022; Sun et al., 2021). Exposure to higher temperatures and heat stress are
also linked to adverse pregnancy and birth outcomes (Baharav et al., 2023; Jiao et al., 2023; Khalili et al.,
2025; Kuehn and McCormick, 2017; USGCRP, 2023; Weeda et al., 2024; Zhang et al., 2023).
TABLE 5.1 Assessment of Change in Evidence of Risks from Climate Change Since EPA (2009)
Assessment of Change in Evidence of Health
Risks from Climate Change Since 2009
New Areas of Evidence
Type of Effect Areas Addressed in EPA (2009) Since EPA (2009)
Exposure to extreme heat
Effects on nutrition
NOTES: Large arrows indicate topics for which the evidence has continued to accumulate. Small arrows indicate
areas for which a potential health effect has been identified and further studies are ongoing.
1
See Weather Related Fatality and Injury Statistics at https://www.weather.gov/hazstat (accessed September 8,
2025).
Prepublication Copy
BOX 5.1
The Role of Adaptation in Reducing Risks to Human Health
Adaptation can help to reduce health risks from climate change. For example, heating and air conditioning
can reduce risks from outdoor temperature extremes. A study of non-accidental death due to hot days found that
this risk declined from 10.6% to 0.9% from the 1960s, associated with increases in air conditioning and
controlling for parameters such as geographic location and mean summer temperature, while noting that “the
remaining 20% of un-air-conditioned housing are not randomly located, but primarily in areas that have less
summer heat, but where summer temperatures are likely to increase” (Nordio et al., 2015).
The potential effectiveness of adaptation measures in reducing future climate-driven health risks is uncertain.
Predicting their success requires assumptions about many factors that are unrelated to climate, including human
behavior, government policies, and technological advances, which are not explored in detail in this report.
Furthermore, there are limits to the ability to adapt to climate-health risks, benefits from available adaptations may
be uneven and incomplete, and available adaptation approaches may differ in utility locally and regionally. In
some instances, effective adaptation approaches are known but remain unimplemented or have been applied
inconsistently. For other climate-related health risks, evidence-based adaptation strategies have yet to be
identified.
Not all groups and communities have the same access to adaptations that mitigate health risks. For example,
communication gaps can limit awareness of increased heat risks and readiness strategies to reduce such risks in
places that have not historically dealt with high temperatures (Healy et al., 2023; Howe et al., 2019). Countries
such as the United States show weaker temperature-mortality links than lesser developed countries, largely due to
the availability of air conditioning, but not all communities have access to this option (Carleton et al., 2022). For
example, outdoor workers exposed to extreme heat and wildfire smoke may not be able to take breaks indoors
with air conditioning or use protective equipment like masks consistently. Finally, adaptation can be costly or
resource intensive. For example, installing home air filtration systems to reduce risks from exposure to poor air
quality or restricting development in the wildland–urban interfaces to reduce risks associated with wildfires,
require financial and policy resources.
Rising temperatures increase risks for workers in sectors such as construction, agriculture,
transportation, warehousing, and waste management (USGCRP, 2023). Other outdoors workers,
including those in landscaping, natural resources management, and firefighting, are also affected by
outdoor temperatures, as are people who live in poorly insulated or unshaded homes and the unhoused.
For example, a 2022 meta-analysis of occupational heat exposure examined 2,409 outdoor workers across
41 jobs in 21 countries, including the United States. The study’s findings suggest that occupational heat
stress elevated workers’ core and skin temperatures, heart rate, and the concentration of dissolved
chemicals and particles in urine (Ioannou et al., 2022). Among workers routinely exposed to heat stress
(≥6 hours/day, 5 days/week, for ≥2 months annually), a study found that approximately 15% developed
kidney disease or acute kidney injury (Flouris et al., 2018). Increased specific humidity in some heat-
prone areas decreases evaporative cooling through sweat and can exacerbate heat stress, although results
of epidemiological studies exploring the role of specific humidity in heat-related health outcomes have
been mixed (Baldwin et al., 2023). With global warming, heat reaches unsafe thresholds for sustained
labor earlier in the morning, making it harder to adapt shifts to safer hours (Parsons et al., 2021). Heat
waves can also reduce the places where fans, rather than air conditioning, are able to provide sufficient
cooling, amplifying the impacts on poorer communities (Parsons et al., 2023).
In the United States, recent studies have assessed excess heat-related deaths attributable to
climate change, identifying impacts on mortality although finding that percentages attributable
specifically to human-induced climate change remain relatively small. As noted for many health effects, a
complex interplay of factors beyond a location’s recorded temperature affects health outcomes. For
example, one study which generalized local epidemiological evidence across the contiguous United States
found 12,000 (95% confidence interval 7,400-16,500) heat-related premature deaths annually in the
United States averaged over 2010-2019 (Shindell et al., 2020). Using data from sites in the United States
Prepublication Copy
and sites around the globe, researchers estimated that 37.0% (20.5-76.3%) of warm-season heat-related
deaths could be attributed to human-induced climate change (34.7% in the United States); this analysis
assessed that a 0.30% increase (95% CI 0.01, 0.76) of heat-related mortality in the United States is
attributable to human-induced climate change (Vicedo-Cabrera et al., 2021). A recent analysis of U.S.
mortality during 1999-2023 in which heat was an underlying or contributing factor observed an increasing
trend; further research will be needed to understand whether such a trend is directly attributable to climate
change (Howard et al., 2024).
EPA (2009) found that “Some reduction in the risk of death related to extreme cold is expected”
as a result of climate change, but “it is not clear whether reduced mortality from cold will be greater or
less than increased heat-related mortality in the United States.” Many factors are involved in evaluating
cold-induced deaths, and cold-related mortality reductions with climate change have not been observed
and remain unclear.
Factors important to assessments of temperature-related morbidity and mortality include not only
the temperature maximum or minimum itself, but also whether the extreme event happens earlier or later
in the season, its duration, how it relates to temperatures normally experienced in that region, the
influence of seasonality on infectious disease transmission, intersections with a person’s age and the
aging population, preexisting health conditions, and other physical and biological factors, including where
they live and work (Healy et al., 2023; Liu et al., 2025). Some parts of the United States may experience
increases in heat-related mortality and other locations might witness reductions in cold deaths (Lee and
Dessler, 2023). On a population level, reductions in cold-related mortality could possibly be expected in a
few regions under climate warming scenarios. Epidemiological studies that examine the number of deaths
associated with cold have found conflicting results. For example, an analysis of U.S. mortality trends
from 1999-2022 reported a 3.4% annual increase in age-adjusted cold-related mortality rates, with a sharp
rise after 2017 (12.1% annually), suggesting that cold-related mortality has not uniformly declined over
time (Liu et al., 2025). Methodological factors contribute to the complexity of assessing relative heat- and
cold-related mortality data, including choices about the temperatures to use and the exposure-response
relationship; for example, Alahmad et al. (2025) have noted that the area under the curve for heat is only a
third of the total area, contributing to calculations of greater numbers of cold-associated death. The
seasonality of cold-related deaths also corresponds to other factors beyond temperature, such as exposure
to influenza, which increases in moderately cold and drier conditions, and to socioeconomic status (Ebi
and Mills, 2013). The existence of multiple factors that follow the same seasonal pattern as temperature
makes attribution of the observed increases in health impacts as temperature reach their coldest seasonal
levels difficult. Evidence since EPA (2009) strengthens association with adverse heat-related health
consequences, while adding nuance around how changes to climate are anticipated to affect deaths from
heat versus cold, suggesting a need for further research in this area to understand the balance of effects.
Studies over the past 15 years have expanded the understanding of how climate change affects
ozone and airborne particulate matter, and how exposure to these air pollutants negatively affects human
health. In 2009, EPA concluded that “[t]he evidence concerning adverse air quality impacts provides
strong and clear support for an endangerment finding” (EPA, 2009a). This conclusion was largely based
on strong evidence that climate change is increasing ground-level (tropospheric) ozone concentrations.
The report cited uncertainty in the directional effect on particulate matter. Since 2009, several
assessments have documented the air quality impacts of climate change, with particular focus on ground-
level ozone and particulate matter (see Section 3.5). Evidence supports the EPA (2009) conclusion on
ground-level ozone and has expanded understanding of the health impacts. For particulate matter,
Prepublication Copy
evidence now points to increases in atmospheric concentrations under climate change in some U.S.
locations, especially in areas prone to wildfires and dust. Because exposure to these pollutants affects a
range of health outcomes, including premature mortality, cardiovascular effects, and respiratory effects,
climate-driven increases in ozone and particulate matter have deleterious health impacts.
Key authoritative assessments published since EPA (2009) conclude that climate change worsens
air pollution. For example, USGCRP (2023) concluded with medium confidence that air quality would
worsen in many parts of the United States, with harm to human health and increased premature death
being very likely (high confidence). “Extreme heat events, which can lead to high concentrations of air
pollution, are projected to increase in severity and frequency (very likely, very high confidence), and the
risk of exposure to airborne dust and wildfire smoke will increase with warmer and drier conditions in
some regions (very likely, high confidence)” (p. 14-5). New research published since this assessment
continues to support this conclusion. The sections below describe the current state of knowledge on ozone
and particulate matter.
Many people are exposed simultaneously to multiple air pollutants, heat, pollen, and other
climate-sensitive risk factors. In 2020, a systematic literature review found sufficient and moderate-
quality evidence for synergistic effects of heat and air pollution (Anenberg et al., 2020). While limited
evidence prevented conclusions from being drawn about synergistic effects from co-exposure to these risk
factors with pollen, the authors concluded that “many disease states, including heart and lung disease,
share a common pathway in which exposure to heat, air pollution, and pollen cause systemic and organ-
specific inflammation and cellular damage.” Since that review was published, additional studies have
found that co-exposure to heat and air pollution had larger effects beyond the sum of their individual
effects (Chen et al., 2024; Rahman et al., 2022; Rai et al., 2023; Stafoggia et al., 2023).
Ground-Level Ozone
EPA (2009) concluded that climate change is expected to worsen ozone pollution across broad
regions of the United States, increasing risks of respiratory illness, premature death, and ecological harm,
even as the effects on particulate matter remain uncertain. Recent literature supports the EPA (2009)
conclusion regarding ground-level (tropospheric) ozone and adds a fuller understanding of the multiple
drivers of ozone increases under climate change as well as harmful interactions with other exposures like
heat and particulate matter. Climate change contributes to increases in ozone exposure on both short-term
and long-term time scales (see Chapter 3).
Ozone exposure is associated with respiratory effects (EPA, 2020; Holm and Balmes, 2022), pre-
term birth (Rappazzo et al., 2021), and premature mortality from all causes and from cardiopulmonary
disease (Jerrett et al., 2009; Lim et al., 2019; Turner et al., 2016). Studies show that ozone concentrations
tend to be higher in areas of the United States that are suburban, exurban, or rural; in wealthier
neighborhoods; and in areas with a higher fraction of the population that is non-Hispanic Asian or non-
Hispanic White (Collins et al., 2022; Liu et al., 2021).
USGCRP (2016) concluded: “Climate change will make it harder for any given regulatory
approach to reduce ground-level ozone pollution in the future as meteorological conditions become
increasingly conducive to forming ozone over most of the United States [Likely, High Confidence].
Unless offset by additional emissions reductions of ozone precursors, these climate-driven increases in
ozone will cause premature deaths, hospital visits, lost school days, and acute respiratory symptoms
[Likely, High Confidence]” (p. 9).
Evidence since 2009 shows that long-term exposure to ozone has health effects beyond those
from short-term exposure to ozone. Epidemiological studies increasingly show health effects associated
with long-term ozone exposure, including chronic respiratory disease and premature mortality among
adults and decreased lung function and lung function growth among children (Di et al., 2017; Hao et al.,
2015; Holm and Balmes 2022; Kazemiparkouhi et al., 2020; Lim et al., 2019; Turner et al., 2016). For
example, a large prospective cohort of U.S. adults with 17 years of follow-up from 1995 to 2011 found
that for each 10 part per billion increase in the annual average 8-hour daily maximum ozone exposure,
Prepublication Copy
ischemic heart disease increased by 6% (95% confidence interval: 2%-9%) and chronic obstructive
pulmonary disease increased by 9% (95% confidence interval: 3%-15%) (Lim et al., 2019).
Methane (CH4) emissions also contribute to long-term ozone concentrations (see Chapter 2
Figure 2.2 for information on annual CH4 emissions and Chapter 3 for discussion of climate effects on air
quality) and to health impacts. An analysis led by the UN Environment Programme reported that every 10
million tonnes of CH4 emissions leads to approximately 430 (approximately 290 to 550) premature
respiratory deaths and approximately 330 (approximately 110 to 540) premature cardiovascular deaths in
the United States attributable to ozone in persons aged 30 and older, along with approximately 150
respiratory hospitalizations and approximately 1500 asthma-related accident and emergency department
visits in the United States due to ozone exposure (UNEP and CCAC, 2021).
Taken together with evidence of conditions favorable to creating more ozone pollution (see
Chapter 3), this evidence supports the EPA (2009) conclusion that GHGs and climate change exacerbate
ozone pollution, adversely affecting health in the United States. Furthermore, the new evidence adds
context to that understanding, including climate-sensitive drivers of ozone pollution, and shows that
health outcomes are amplified by climate-driven increases in co-occurrence of ozone exposure with heat
and particulate matter exposure, and by synergistic health effects from these multiple exposures.
Particulate Matter
EPA (2009) found that “the directional effect of climate change on ambient particulate matter
levels remains uncertain.” Since 2009, evidence has shown that climate change is altering particulate
matter levels across the United States, with climate-driven increases in the Western United States due to
increased wildfire smoke and dust.
Fine particulate matter, or PM2.5, referring to particles that are 2.5 microns or smaller in diameter,
are solid and liquid particles that can penetrate deeply into the human lung and affect a wide range of
biological systems. PM2.5 is a mixture of chemical components that varies geographically, mainly driven
by local emission sources and atmospheric transport of air pollution regionally. Major components
include black carbon, organic carbon, nitrate, sulfate, ammonium, and metals and other trace elements.
The EPA has found PM2.5 to be causally associated with cardiovascular effects and mortality and likely to
be causally associated with respiratory effects, nervous system effects, and cancer (EPA, 2022). Many
studies show that PM2.5 is inequitably distributed, with communities with lower income and higher
proportions of non-White populations most exposed (Colmer et al., 2020; Jbaily et al., 2022; Ma et al.,
2023b).
USGCRP (2016) found that climate change is expected to alter several meteorological factors that
affect PM2.5. Factors that are expected to increase PM2.5 include increased humidity, increased stagnation
events, and increased biogenic emissions. Factors that are expected to decrease PM2.5 include increases in
precipitation and enhanced atmospheric mixing. The USGCRP report also found links between climate
change and increased frequency and length of wildfires and wildfire seasons, with associated emissions
and harmful impacts on health. It found that “wildfires may dominate summertime PM2.5 concentrations,
offsetting even large reductions in anthropogenic PM2.5 emissions” (p. 77). Climate change has been
projected to increase drought in some regions, which can also lead to more airborne dust exposure.
USGCRP (2023) similarly found that most known climate-related drivers of PM2.5 increase
concentrations, including wildfires, heat waves, temperature, drought, biogenic emissions, and humidity,
and that more precipitation would lower PM2.5 levels, as precipitation is a main removal mechanism of
PM2.5 from the air. However, to date there is no consensus on the impacts of climate-driven changes in
regional transport and stagnation on PM2.5.
Wildfires
EPA (2009) found that “In some regions, changes in the mean and variability of temperature and
precipitation are projected to increase the size and severity of fire events, including in parts of the United
Prepublication Copy
States” (p. 86), noting that “Increase in wildfire frequency associated with a warmer climate has the
potential to increase PM levels in certain regions” (p. 94). While EPA (2009) addressed impacts of
wildfires on welfare, it did not expand on the health impacts of wildfires in detail. Since 2009, a large
body of literature has developed on climate change-driven variations in wildfire smoke exposure and
associated health impacts in the United States. See Chapter 3 for a discussion of wildfire trends and
variability, including an increase in burned acreage in the West.
The presence of more wildfires across the U.S. landscape and wildfire smoke in the skies is one
condition that has affected the United States with more severity and scope than expected since 2009.
Wildfires threaten health directly—through injuries, burns, and heat exposure—and indirectly, through
smoke-related respiratory illness and trauma-related mental health harms (Gould et al., 2024; Lei et al.,
2024; Ma et al., 2023a).
These health impacts are felt in communities directly affected by the fire and first responders
such as wildfire fighters, as well as in distant locations as smoke can travel across state and even national
boundaries. Wildfire smoke contributes to both short-term health outcomes from acute exposures and
chronic health outcomes from long-term average PM2.5 concentrations. Wildfires affect air quality as
extreme episodic events and as contributors to everyday exposures (Gould et al., 2024).
USGCRP (2016) concluded that: “Wildfires emit fine particles and ozone precursors that in turn
increase the risk of premature death and adverse chronic and acute cardiovascular and respiratory health
outcomes (Likely, High Confidence). Climate change is projected to increase the number and severity of
naturally occurring wildfires in parts of the United States, increasing emissions of particulate matter and
ozone precursors and resulting in additional adverse health outcomes (Likely, High Confidence)” (p. 85).
As the literature continued to build, the USGCRP (2023) concurred with the 2016 report that climate
change contributes to more frequent and severe wildfires that worsen air quality in many regions, while
noting that “Although large challenges remain, new communication and mitigation measures are reducing
a portion of the dangers of wildfire smoke (medium confidence)” (p. 14-9).
Wildfires release a complex mix of particulate matter, carbon monoxide, nitrogen oxides, and
volatile organic compounds that can harm human health. Wildfire-specific particulate matter likely has a
different chemical profile compared with particulate matter originating from other sources and exposure
to PM2.5 from wildfire smoke has been reported to be more harmful to health than PM2.5 from other types
of sources although more research is needed to understand this differential toxicity (Aguilera et al., 2021;
Alari et al., 2025; Gould et al., 2024). Wildfire smoke from fires in urban and industrial areas and the
wildland-urban interface can also contain toxic metals such as lead and mercury, plasticizers, and other
pollutants (NASEM, 2022). Wildfire smoke has also been documented to contain carcinogens such as
benzene, benzo[a]pyrene, hexavalent chromium, and dibenz[a,h]anthracene (Naeher et al., 2007). As a
recent review noted, “The amount and composition of pollution emitted from a specific fire vary
depending on the fire’s size, temperature of combustion, materials burned (e.g., grasses, tree species,
buildings, vehicles), distance the smoke has traveled, and environmental conditions like wind speed,
temperature, and humidity” (Gould et al., 2024, p. 279; Montrose et al., 2022).Consistent with these prior
assessments, new studies find that wildfires are offsetting reductions in anthropogenic PM2.5 emissions
over the Western United States and increasing daily cumulative smoke PM2.5 exposure for the average
person in the United States compared with the 2006-2019 average (Lancet Countdown, 2024; Wei et al.,
2023). The number of people in the United States that have experienced at least 1 day with wildfire
smoke of PM 2.5 > 100 μg m−3 has significantly increased, with about 25 million people exposed in 2020
(Childs et al., 2022).
Observational studies indicate that exposure to wildfire smoke, like exposure to air pollution from
other sources, is associated with a spectrum of adverse health outcomes. Substantial literature documents
wildfire smoke’s impacts on respiratory health (Zhou et al., 2021). Gould et al. (2024) found that “same-
day all-cause mortality increased by 0.15% [95% confidence interval (CI) 0.01–0.28%] per 1-μg m–3
increase in wildfire-specific PM2.5. There were robust positive associations between wildfire PM2.5 and
same-day respiratory outcomes: Respiratory hospitalizations increased by 0.25% (95% CI 0.09–0.52%)
and respiratory ED visits increased by 0.36% (95% CI 0.19–0.53%) per additional 1-μg m–3 increase in
Prepublication Copy
ambient wildfire smoke PM2.5. [The study] found a non–statistically significant 0.06% (95% CI 0.00–
0.12%) increase in same-day cardiovascular hospitalizations and no meaningful change in same-day
cardiovascular ED visits (–0.03%; 95% CI–0.18–0.12%) per additional 1-μg m–3 increase in ambient
wildfire smoke PM2.5. For all outcomes except respiratory hospitalizations and cardiovascular ED visits,
there was evidence of heterogeneity in effects across studies (i.e., Q-statistic p < 0.05)” (p. 282).
Research is also associating exposure to wildfires with negative pregnancy and birth outcomes,
such as preterm birth, potentially through the effects of both maternal stress from experiencing the
wildfire and exposure to wildfire smoke (Heft-Neal et al., 2021). Exposure to wildfire smoke has also
been associated in recent literature with declines in mental health and to worsened cognitive outcomes
(Eisenman and Galway, 2022; Xu et al., 2020). Wildfire smoke can spread spores and microbes that
contribute to fungal disease and anti-microbial resistance (Mulliken et al., 2023; Salazar-Hamm and
Torrez-Cruz, 2024). New research also indicates that wildfire smoke exposure may contribute to skin
diseases, eye conditions, and cancer.
One study estimated that of 164,000 estimated wildfire PM2.5-related deaths over 15 years in the
United States, about 10% (or 15,000) were linked with climate change-driven increases in wildfire. These
premature deaths from climate change-driven increases in wildfire smoke translated to a cumulative
economic burden of $160 billion. More than one-third of the climate change-driven PM2.5 deaths occurred
in 2020, with monetized damages of $58 billion (Law et al., 2025). Another study estimated that long-
term exposure to carbonaceous PM2.5 from fire smoke led to 7,455 (95% CI: 6,058, 8,852) premature
deaths across the continental United States each year, with monetized damages of $68.3 billion (95% CI:
$31.9 billion, $104.0 billion) (Jin et al., 2025).
Dust
EPA (2009) found that “[particulate matter] and [particulate matter] precursor emissions are
affected by climate change through physical response (windblown dust), biological response (forest fires
and vegetation type/distribution), and human response (energy generation). Most natural aerosol sources
are controlled by climatic parameters like wind, moisture, and temperature; thus, human-induced climate
change is expected to affect the natural aerosol burden” (p. 94). Since 2009, evidence linking climate
change with increased airborne soil dust and associated health effects has continued to build, particularly
for areas that are warmer and drier, such as the Southwest United States (see section on Climate Sensitive
Diseases).
Dust is a component of particulate matter in the air and is a key contributor to PM10, or particles
10 microns or smaller in diameter (NASEM, 2025). Major emission sources include soil entrainment into
the air and anthropogenic activities, such as road and vehicle tire wear. Climate change in other world
regions can impact dust concentrations over the United States, as dust from African deserts reaches the
Caribbean and dust from Asia reaches the Western United States and Hawaii (Rosas et al., 2025; Yu et
al., 2019). Dust exposure can lead to a variety of deleterious health outcomes, including premature death,
allergies, asthma attacks, and respiratory infection. A large scoping review found that among 204
epidemiological studies, over 80% reported positive associations between dust and adverse health
outcomes (Lwin et al., 2023).
USGCRP (2016) concurred with the EPA (2009) statement, finding that “climate-driven changes
in meteorology can also lead to changes in naturally occurring emissions that influence air quality (for
example, wildfires, wind-blown dust, and emissions from vegetation)” (p. 71). USGCRP (2023) also
concluded in Key Message 14.1 that “the risk of exposure to airborne dust and wildfire smoke will
increase with warmer and drier conditions in some regions (very likely, high confidence).” These findings
for the United States are consistent with the 2024 Lancet Countdown report that has a global scope. As
part of that effort, the Lancet Countdown found that “The hotter and drier weather conditions are
increasingly favoring the occurrence of sand and dust storms” (Romanello et al., 2024). Recent studies
build on earlier findings around how climate change affects dust concentrations in parts of the United
States and provide more information about contributions of different climate-sensitive drivers. In a
Prepublication Copy
historical, observational study, Achakulwisut et al. (2018) found that drought and soil dryness were key
drivers of increased fine dust over the Southwest United States, with 0.22-0.43 μg/m3 increase in fine
dust for each unit increase in the 2-month Standardized Precipitation-Evapotranspiration Index, an
indicator of soil dryness.
Climate change can also lead to more entrainment of dust from exposed lakebeds into the air
(NASEM, 2020, 2025; West et al., 2023). Changing precipitation patterns, higher temperatures, persistent
droughts, less water inflow from reduced snowpack, and increased evaporation, among other climate-
sensitive conditions, can result in lower water levels for lakes in parts of the United States, such as the
Great Salt Lake (Baxter and Butler, 2020). With a larger area of exposed lakebed, more dust can become
entrained into the air, exposing people in nearby communities and across a broader area (Grineski et al.,
2024). Lakebed dust often contains metals, pathogens, and other health-harmful agents (Putman et al.,
2025). More exposure to lakebed dust could result in a variety of health outcomes, with potentially higher
risks for children (Putman et al., 2025). Evidence on the relative contributions of climate change and
water management practices to declining lake levels is limited and differs across lakes.
Indoor air
Impacts of climate change on indoor air quality were not directly addressed in EPA (2009).
Indoor air quality is affected by outdoor air coming in, contaminants generated indoors (e.g. mold, dust
mites, volatile organic compounds and other chemicals off gassing from building materials, indoor
combustion), indoor temperatures, and other factors (NASEM, 2024c). As Americans spend most of their
time indoors (Klepeis et al., 2001), changes in indoor air quality can have important effects on public
health.
UGSCRP (2016) identified climate impacts on indoor air quality as an emerging issue. The report
highlighted multiple pathways through which climate change can negatively affect indoor air quality,
including worsening outdoor air pollution that infiltrates indoors; altered patterns of indoor-outdoor air
exchange; and more favorable conditions for growth and spread of pests, infectious agents, and disease
vectors. Evidence showing negative impacts of climate change on indoor air quality has continued to
build since the 2016 report, particularly related to indoor air quality during wildfire smoke events, mold
driven by building dampness, and climate sensitivity of pollutants originating indoors.
Indoor exposure to wildfire smoke is a concern, especially as common public health messaging
during wildfire smoke events is to stay indoors. USGCRP (2023) found that “Research investigating
indoor concentrations during wildfire smoke events is preliminary, and there is a specific need to
understand how indoor concentrations vary between socioeconomic groups during wildfire smoke events”
(p. 14-23). Recent studies show that volatile organic compounds from wildfire smoke can persist indoors
for days after the smoke event (Dresser et al., 2024; Li et al., 2023), and that smoke and other forms of
outdoor air pollution increase indoor PM2.5 levels, particularly in lower income areas (Krebs and Neidell,
2024). Wildfire smoke exposure and increased indoor crowding to avoid outdoor smoke is also associated
with increased risk of infection from viruses (Arregui-García et al., 2025; Mahendran et al., 2025; Orr et
al., 2025), such as influenza and COVID-19.
Climate change can also influence indoor air quality from pollutants originating indoors. Rain,
flooding, and humidity changes affect building dampness, leading to mold and other microbial agents that
increase risk of allergic rhinitis, asthma, and other respiratory conditions (Eguiluz-Gracia et al., 2020;
WHO, 2009). Increased temperature and humidity can alter rates of chemical offgassing from building
and furniture materials, as well as chemical reaction rates, both of which influence levels of indoor air
pollutants (Abbat and Wang, 2020; Salthammer and Morrison, 2022).
These potential impacts of climate change on indoor air quality likely differ across households
and other buildings, driven by geographic and building-specific factors such as building codes, building
materials, presence of air filtration devices, and climate control, making it challenging to evaluate the
combined impact of the effects described here in different locations (Mansouri et al., 2022).
Prepublication Copy
Fossil fuel combustion leads to emissions of both carbon dioxide (CO2) and co-emitted
particulate matter and ozone and particulate precursors that directly affect air quality (NASEM, 2024;
USGCRP, 2023). Information about air quality and health co-benefits from reduced fossil fuel
combustion was building in 2009 (Bell et al., 2008), and since that time has expanded with many
quantified and monetized estimates of co-benefits from reducing fossil fuel combustion in power,
residential, transportation, and other sectors (Balbus et al. 2015; Buonocore et al., 2016; Garcia et al.,
2023; Levy et al., 2016; Sergi et al., 2020). For example, a review of the health impacts during the first 6
years of the Regional Greenhouse Gas Initiative—a policy that reduced GHG emissions in the
Northeastern and Atlantic regions of the United States—found that “These benefits include hundreds of
avoided cases of premature deaths, heart attacks, asthma attacks, and hospital admissions, and tens of
thousands of avoided cases of other health symptoms, lost work days, and restricted activities” (Manion et
al., 2017, p. 38). Several studies project that future economic benefits associated with avoiding health
effects of co-emitted pollutants are substantial (e.g., McDuffie et al., 2023; Shindell et al., 2024; West et
al., 2023).
Extreme weather events contribute to injury and illness, exacerbate chronic disease, and
affect mental health (Ebi et al., 2021; USGCRP, 2016). The occurrence of extreme weather events
can also disrupt critical infrastructure and health care systems, reducing access to care, disrupting
supply chains, and contributing to mortality (Salas et al., 2024).
Chapter 3 discusses the ways that climate change is affecting extreme weather. This section
addresses health effects associated with extreme weather events, noting that it is challenging to attribute
health impacts from individual weather events to climate change. Determining how climate change
influences any single weather or climate event requires accounting for multiple natural and human factors
(NASEM, 2016), and the health effects associated with any event also can be affected by multiple factors.
Droughts
Droughts are projected to become more frequent, longer lasting, and more severe across some
regions of the United States (Martin et al., 2020; Overpeck and Udall., 2020; Tripathy et al., 2023).
Research has linked drought to health consequences including higher risks of respiratory, cardiovascular,
and all-cause mortality in several U.S. regions, particularly among older adults, women, and rural
residents (Abadi et al., 2022; Gwon et al., 2023, 2024, 2025; Salvador et al., 2023). Drought compounds
the health risks of other extreme events such as heat waves and dust storms (Leeper et al., 2025) and
combined drought–heat events have been associated with increased mortality in people with chronic lung
disease (Rau et al., 2025). Beyond physical health, drought contributes to mental and occupational health
risks, with studies showing greater stress among farmers during drought years (Berman et al., 2021).
Climate change and drought also impact water quality (see discussion in Chapter 6).
Hurricanes
Globally, the share of hurricanes reaching the most intense categories has increased over the
past four decades, and although landfalls in the United States have not increased, there is emerging
evidence that U.S. hurricanes are moving more slowly at landfall, producing heavy rainfall, damaging
wind, and coastal flooding (see discussion in Chapter 3). Several studies have analyzed the mortality
associated with hurricanes. For example, an analysis of two approaches found that Hurricane Maria
was responsible for 1,191 excess deaths using census population data, and 2,975 excess deaths (95%
CI 2,658-3,290) from September 2017 to February 2018 when accounting for demographic shifts that
Prepublication Copy
had also taken place over that time (Santos-Burgoa et al., 2018). Effects can persist for a prolonged
period after the event itself; a modeling study examining longer-term, indirect effects of tropical
cyclones on mortality in the United States estimated that the average tropical storm contributed 7,000
to 11,000 excess deaths (Young and Hsiang, 2024). The study also examined 501 historical storms
between 1930 and 2015 and estimated a tropical cyclone-related mortality burden of 3.2-5.1% of all
deaths in the Atlantic coastal region between 1930 and 2015 (Young and Hsiang, 2024). Evidence
shows an increase in the average number of deaths per tropical cyclone in the United States since
2001, due to a combination of storm factors, shifting of population spatial distribution towards coastal
areas, and demographic trends. Young and Hsiang (2024) found no evidence of adaptation reducing
the deadliness of these storms.
Floods
Floods are associated with human health impacts, as described below. Many non-climate factors
contribute to the risks to human health from floods, such as emergency preparedness and response and the
location and state of infrastructure. Changes in heavy precipitation and sea level rise associated with
climate change (see Chapter 3) may also contribute to the risk. A recent global analysis of flood fatalities
by Jonkman et al. (2024) found no trend in flood-related mortality has been observed.
Extreme rainfall and flooding have been linked with hospital admissions and adverse health
outcomes, including increased risk of injury, infectious diseases, increased morbidity and mortality from
cardiovascular disease and other causes (Aggarwal et al., 2025; He et al., 2024; Lynch et al., 2025;
Wettstein et al., 2025). Floods can also damage or impede access to critical infrastructure including
hospitals, disrupt medical supply chains, and cut people off from care (Wu et al., 2024). For example, a
Veterans Affairs hospital closed for six months after Hurricane Sandy, and the study found that the
“temporary period of decreased access to health care services was associated with increased rates of
uncontrolled hypertension, but not with increased rates of uncontrolled diabetes or hyperlipidemia, more
than 1 year after the Manhattan VA facility reopened” (Baum et al., 2019). More than 700 hazardous
waste sites are located in high-risk flood zones, increasing concerns about exposure to toxic chemicals
(GAO, 2019). Beyond physical illness and death, exposure to floods can contribute to adverse pregnancy
outcomes and leave lasting mental health impacts, especially for children (Wu et al., 2024).
EPA (2009) noted that many human diseases were sensitive to weather and the USGCRP (2016)
report on climate and health subsequently stated with high confidence levels that climate change is
harming human health by increasing morbidity and mortality. As discussed above, exposure to increased
heat and worsened air quality contributes to negative health outcomes. The effects of changing climate on
the distribution of vector-borne diseases effects are briefly described in this section. The effects of
exposure to changing temperatures, weather events, increased PM2.5, and allergens on a variety of chronic
and non-communicable diseases are also briefly discussed.
Vector-Borne Diseases
Climate suitability for various climate-sensitive pathogens and disease vectors has increased since
EPA (2009). Ticks can carry many diseases, including alpha-gal syndrome, Lyme disease, Babesiosis,
Rocky Mountain spotted fever, and others. An increase in the geographic distribution of tick-borne
diseases, anticipated in the 2009 report, has been observed and attributed to climate warming. Persistently
warming temperatures may not only expand their geographic range but also extend their active season
(USGCRP, 2016). One example is the lone star tick, which carries alpha-gal syndrome (inducing meat
allergy) and has dramatically increased its U.S. distribution due to warming (Molaei et al., 2019).
Similarly, Lyme disease—another tick-borne disease described in the 2009 EPA report—has expanded its
Prepublication Copy
range and activity due to climate warming with expansion in the northern United States and decreased
southern activity, recognizing also that multiple factors interact to drive disease transmission (Couper et
al., 2021; Kugeler et al., 2015; USGCRP, 2016). Lyme disease case report maps from the Centers for
Disease Control and Prevention show range expansion from 1995 to 2023 2 with Lyme disease cases
increasing from approximately 23,000 in 2002 to more than 62,000 in 2022 (CDC, 2004; Kugeler et al.,
2022). Couper et al. (2021) examined Lyme disease incidence and found that the clearest climate–Lyme
disease signal was observed in the Northeast United States, where rising annual temperatures were
associated with higher incidence and a high-emissions scenario projected that there could be an estimated
23,619 ± 21,607 additional cases by 2050; however, this projection had uncertainty and no significant
increases in disease incidence were projected for other regions. Couper et al. (2021) also found no
significant increases or decreases in Lyme disease across any region under a moderate-emissions
scenario, underscoring the multi-factor and regionally variable relationship between climate change and
Lyme disease dynamics.
Dengue, a mosquito-borne viral disease, has increased in geographic range since EPA (2009),
now appearing in non-travelers in Texas, Florida, Arizona, California, and Hawaii. 3 This shift coincides
with increased mosquito vector activity due to climate warming (Ebi and Nealon, 2016). With rising
temperatures, the aggressive Asian tiger mosquito (Aedes albopictus) has expanded in the United States,
raising concerns about potential outbreaks of diseases from Chikungunya and Zika viruses (Rochlin et al.,
2013). Mordecai et al. (2019) has described how temperature affects both mosquito biting behavior and
transmission of pathogens. West Nile disease, which is also mosquito-borne, has been seen in the United
States since the 2009 report, related to changes in precipitation and heat (Hahn et al., 2015).
Fungal Diseases
Fungal infection is expanding across the United States, particularly in the Western states (Salazar-
Hamm and Torrez-Cruz, 2024). Soil dust contains a variety of microorganisms, including bacteria and
fungi. Recent studies show association between airborne soil dust or dust storms and Valley Fever
(Howard et al., 2024a; Tong et al., 2017). Valley Fever, or coccidioidomycosis, is a fungal infection
resulting from breathing Coccidioides fungal spores that can cause fever, cough, fatigue, shortness of
breath, and other symptoms; its incidence has increased from approximately 2,000 cases in 1998 to
21,000 cases in 2023 4 and its area of endemicity has expanded to include 12 states: Arizona, California,
Colorado, Idaho, Kansas, Nebraska, Nevada, New Mexico, Oklahoma, Texas, Utah, and Washington
(Gorris et al., 2019). In addition, changing weather patterns can affect Coccidioides growth and dispersal,
as the fungus grows in the soil after heavy rainfall and disperses into the area in subsequent hot and dry
conditions (Head et al., 2022).
Evidence also points to changes in geographical extent of histoplasmosis, the most frequent
fungal respiratory infection in the United States The causative agent, Histoplasma capsulatum, is a
dimorphic soil-based fungus endemic to the Midwest United States, Latin America, Africa, South Asia,
and the Caribbean. Approximately 60-90% of people living in areas surrounding the Ohio and Mississippi
river valleys have been exposed to Histoplasma. 5 Symptoms include fever, chills, headache, muscle
aches, fatigue, and cough. 6 Histoplasmosis affects people who are immunosuppressed more severely.
While histoplasmosis is far less studied than Valley Fever, the area of endemicity in the United States is
2
See https://www.cdc.gov/lyme/data-research/facts-stats/lyme-disease-case-map.html (accessed September
8, 2025).
3
See https://www.cdc.gov/dengue/outbreaks/2024/index.html.
4
See Reported Cases of Valley Fever at https://www.cdc.gov/valley-fever/php/statistics/index.html (accessed
September 8, 2025).
5
See https://www.cdc.gov/histoplasmosis/php/statistics/?CDC_AAref_Val=https://www.cdc.gov/fungal/diseases/
histoplasmosis/statistics.html.
6
See https://www.lung.org/lung-health-diseases/lung-disease-lookup/histoplasmosis/symptoms-diagnosis.
Prepublication Copy
spreading northwest, to Minnesota, Wisconsin, Michigan, Montana, and Nebraska (Hepler et al., 2022;
Maiga et al., 2018).
Antimicrobial Resistance
EPA (2009) did not address antimicrobial resistance as a climate-sensitive issue. Antimicrobial
resistance causes significant global morbidity, with projections estimating 8.2 million annual deaths
associated with antimicrobial resistance by 2050 (Naghavi et al., 2024). Several studies show that higher
temperatures accelerate bacterial growth, mutation rates and gene transfer, increasing resistance risk to
antibiotics (McGough et al., 2020; Van Eldjik et al., 2024). Heat waves and droughts concentrate
antibiotics and resistant bacteria in water systems, enhancing opportunities for developing resistance,
while extreme rainfall spreads antibiotic resistant pathogens from sewage and farms where antibiotics are
used for animal health (MacFadden et al., 2018). Better understanding this area and its potential
intersections with climate change continues to be a topic of exploration.
Waterborne Diseases
Heavy precipitation and drought are linked to waterborne disease outbreaks. Excessive rainfall
mobilizes pathogens into water supplies, while droughts disrupt sanitation. Contrary to the 2009 report’s
claim that flood-related infectious disease risks are low in high-income countries (EPA, 2009), recent
studies show heavy precipitation increases gastrointestinal illnesses in the United States (De Roos et al.,
2020; Haley et al., 2024). Hurricanes and extreme weather events can also introduce pathogens into water
systems via disrupted sanitation infrastructure. For example, Vibrio parahaemolyticus is a waterborne
bacterium that causes seafood-associated diarrheal disease in the United States, while Vibrio vulnificus is
also found in marine settings and causes serious wound infections and diarrheal illness. Increasing water
temperatures and other changes to coastal waters caused by climate are predicted to enhance Vibrio
replication with resultant increased infection from contaminated shellfish or wound exposure to
contaminated water (Hayden et al., 2023; Schets et al., 2025; USGCRP, 2016). Extension northward
along the East Coast and increases in the numbers of reported non-foodborne cases of Vibrio vulnificus
since 1988 have been observed (Archer et al., 2023; Brumfield et al., 2025).
Moreover, as water temperatures rise and more humans use recreational water due to heat,
increasing infections with Naegleria fowleri, a thermophilic ameba which causes meningoencephalitis, is
possible as well as a northward expansion of this disease (Heilmann et al., 2024; USGCRP, 2016). The
potential for exposure to toxins associated with harmful algal blooms is another potential impact on health
(see Chapter 6).
Non-Communicable Diseases
Exposure to heat, ozone, particulate matter, and extreme events also impact several non-
communicable and chronic diseases associated with the cardiovascular, renal, and pulmonary systems.
The potential impacts extend to psychological and mental health and nutrition, areas not addressed in
EPA (2009).
Cardiovascular health: Cardiovascular diseases are the world’s leading cause of disability and death. In
2022, 941,652 U.S. deaths were attributable to cardiovascular disease for all ages (Martin et al., 2025).
Pollution has had a major impact on cardiovascular morbidity and mortality. Short-term variation in PM2.5
levels (from hours to days) is associated with increased risks of myocardial infarction, stroke, and death
from cardiovascular disease (Rajagopalan and Landrigan, 2021). During heat waves, a meta-analysis of
266 papers found the risk of cardiovascular disease-related mortality increased by 11.7% (95% CI 9.3-
14.1%) with the risk increasing as heat wave intensity increased (Liu et al., 2022a). Both hot and cold
temperature extremes increase risk; an analysis of cardiovascular-related deaths across 27 countries
Prepublication Copy
including the United States found that “hot days (above 97.5th percentile) and cold days (below 2.5th
percentile) accounted for 2.2 (95% empirical CI [eCI], 2.1-2.3) and 9.1 (95% eCI, 8.9-9.2) excess deaths
for every 1000 cardiovascular deaths, respectively” (Alahmad et al., 2023). There are several mechanisms
for how extreme temperature impacts the cardiovascular system. Heat, for example, strains the
cardiovascular system (e.g., dehydration, hypotension, tachycardia, electrolyte shifts). People with heart
failure, coronary disease, or arrhythmia are at highest risk; certain drugs (e.g., β-blockers, antiplatelets)
can impair heat loss or increase heat-myocardial infarction (MI) risk (Chen et al., 2022). The risk has
often been found to be higher in susceptible subgroups, including older people, people with preexisting
conditions, and those who are socioeconomically disadvantaged (Singh et al., 2024). Heat waves and
factors such as ground-level ozone that worsen air quality are associated with higher rates of
cardiovascular mortality (Kazi et al., 2024). Exposure to PM2.5 components, emissions from fossil fuels,
and chemical combustion by anthropogenic sources (e.g., gas stations) are associated with increased
hypertension (Chen et al., 2025; Xu et al., 2022), another risk factor for cardiovascular morbidity. It has
been calculated that for every increment of 10 μg m–3 in PM2.5 the risk of myocardial infarction, stroke, or
cardiovascular-related death increases 0.1-1%, and air pollution is estimated to contribute to 14% of all
stroke-associated deaths (Rajagopalan and Landrigan, 2021; Verhoeven et al., 2021). A meta-analysis of
35 studies showed “increases in particulate matter concentration were associated with heart failure
hospitalisation or death (PM2.5 2.12% per 10 μg/m3, 95% Confidence Interval 1.42-2.82; PM10 1.63% per
10 μg/m3, 95% Confidence Interval 1.20-2.07). Strongest associations were seen on the day of exposure,
with more persistent effects for PM2.5. In the USA, we estimate that a mean reduction in PM2.5 of 3.9
μg/m3 would prevent 7978 heart failure hospitalisations and save a third of a billion U.S. dollars a year”
(Shah et al., 2013, p. 1039). Air pollution has also been associated with an increased risk of atrial
fibrillation and ventricular arrhythmias (Peralta et al., 2020).
Renal health: Extreme temperatures can increase the risk of heat stroke and dehydration leading to heat
exhaustion, hypotension, and acute kidney disease (Glaser et al., 2016). There has also been recent
increase in chronic kidney disease of unknown etiology in rural communities, particularly in outdoor
working farmworkers, in different parts of the worldregions with increased temperatures and decreased
rainfall (Glaser et al., 2016). Current hypothesis suggests that heat along with occupational exposures
plays a role in this current epidemic of unexplained chronic kidney disease in MesoAmerica and Sri
Lanka (Hasson et al., 2023; Johnson et al., 2019; Wijkström et al., 2013). Studies are now ongoing
looking for the cause of unexplained increased renal disease in agricultural workers in the central valley
of California, with a hypothesis that extreme heat is an underlying factor (Bragg-Gresham et al., 2020).
Pulmonary health: Climate change is associated with increased risk factors for respiratory health. This
linkage was noted in EPA (2009); however, the extent of impact was not fully explored. Air pollutants
and extreme temperatures have been associated with an increased risk of chronic obstructive pulmonary
disease (COPD) and asthma exacerbations and higher mortality rates from these respiratory diseases
(Almetwally et al., 2020). Various components of air pollution have been associated with worsened lung
function, asthma exacerbations, and COPD, including PM2.5, nitrogen dioxide, and tropospheric ozone
(Vongelis et al., 2025). Wildfire smoke—particularly PM2.5—is strongly associated with increased
emergency department visits and hospitalizations for respiratory illnesses such as asthma (McArdle et al.,
2023). Particularly susceptible populations include children, adults over the age of 65, pregnant women,
and people in areas where access to climate adaptation and health care may be compromised (Covert et
al., 2023).
Psychological and mental health: The climate’s impact on mental health was not addressed in EPA
(2009). Extreme weather events have been associated with increased rates of anxiety, depression, and
mental health disorders (Barbani, 2025). Air pollution (PM2.5) has been associated with increased
prevalence of mood and psychotic disorders (e.g., schizophrenia) and suicide (Hoare, 2019; Kim, 2018;
Newbury et al., 2019). Heat waves and high temperatures are associated with increased cases of suicide,
Prepublication Copy
hospital visits for mental health issues, crime and violence (Choi et al., 2024; Mahendran et al., 2021;
Nori-Sarma et al., 2022; Thompson et al., 2023). Sleep (particularly deep sleep) is also affected by
ambient temperature and humidity, which may be impacted by climate warming, and can further
exacerbate mental health declines (Li et al., 2025; Okamoto-Mizuno and Mizuno, 2012). There is also
growing evidence that long-term exposure to air pollution may play a role in dementia (Abolhasani et al.,
2023; Best Rogowski et al., 2025). Studies have shown increased anxiety among children and youth
around climate change. A global study of 10,000 youth and young adults in 10 countries found that 59%
reported being extremely worried about climate change, and more than 50% reported negative emotions
such as sadness, anxiety, anger, helplessness, and powerlessness (Hickman et al., 2021). Lastly, one
should note the many psychological impacts from events that displace people from their homes (Bellizzi
et al., 2023).
Nutrition and food safety: Impacts on nutrition were also not addressed in EPA (2009). There are
several pathways by which climate change impacts nutrition and food safety. The rise in temperatures and
variability in precipitation amount and intensity have negatively affected agricultural production (see
Chapter 6 for more detailed discussion of food production and agriculture). Such events strain the global
food systems (Romanello et al., 2021) and can affect crop frequency (number of production seasons per
year) and caloric yields (crop yield in calories produced per acre). Food safety is also threatened by
increased warming. The IPCC 2022 report noted that increasing temperature accelerated the growth of
foodborne pathogens like Salmonella and Campylobacter. Warmer conditions extend the survival of
pathogens in soil, and water and food supply. Alterations to marine ecosystems can disrupt marine food
supplies in some communities (USGCRP, 2023). Elevated atmospheric CO2 concentration is associated
with lower nutritional value in staple crops. Myers et al. tested several crops from different countries
including the United States and found a significant decrease in the concentration of iron and zinc in C3
(cool season) grasses and legumes (Myers et al., 2014). Zinc and iron deficiencies increase the risk of
infections, diarrhea, and anemia. These sequelae would be more impactful in low resourced settings
globally or in malnourished children in the United States. Although CO2 concentrations can increase plant
growth with potential effects on crop yields (see discussion in Chapter 6) the IPCC 2023 Synthesis Report
noted that these benefits are often offset by climate related stresses such as heat, drought and nutrient
limitations.
As reported in EPA (2009), multiple factors influence allergen levels and health consequences
such as allergies and asthma, including changes to CO2 and climate that affect plant growth, distribution,
and allergenicity. Evidence since 2009 continues to indicate that GHGs and associated climate changes
affect airborne allergens in ways that can contribute to allergies and asthma, while recognizing that
development of such conditions depends not only on environmental exposures but also on individual and
genetic factors (Dharmage et al., 2019).
USGCRP (2023) reported that factors such as rising temperatures, changes in precipitation,
elevated CO₂, and higher ozone levels are affecting pollen, with effects than can include extending pollen
seasons, boosting pollen levels, and broadening the geographic range of allergenic plants, while
enhancing pollen allergenicity (Agache et al., 2024; Anderegg et al., 2021; Epstein et al., 2025; Lee et al.,
2023; Paudel et al., 2021; USGCRP, 2023; WHO, 2018; Zhang and Steiner, 2022). For instance, ragweed
pollen season in parts of the United States grew longer by as much as 13 to 27 days between 1995 and
2009 (Ziska et al., 2011). Increases in rainfall and flooding associated with climate change can also foster
mold growth and facilitate the introduction of new allergenic species (Epstein et al., 2025). Extreme
weather can worsen respiratory risks. Thunderstorms can rupture pollen grains, leading to “thunderstorm
asthma,” associated with increases in asthma events (Agache et al., 2024; Beggs, 2024; D’Amato et al.,
2016; Mampage et al., 2022), with a recent analysis finding a 24% increased risk (95% CI 13-36%)
(Makrufardi et al., 2023). Post-disaster studies of hurricanes have found increased mold and endotoxin
Prepublication Copy
exposures and mold reactivity, especially among those with asthma (Chew et al., 2006; Rao et al., 2007;
Sampath et al., 2023). Wildfire smoke, dust, and sandstorms can alter pollen structure, increasing
allergenicity and inflammation and amplifying allergic and respiratory diseases (WHO, 2025). Drier,
hotter conditions may intensify airway inflammation, while air pollution increases allergic inflammation
and susceptibility to viral infections (Burbank, 2025; Edwards et al., 2025; Wright and Demain, 2024).
Elevated atmospheric CO2 concentration leads to more vigorous growth in many plant species,
which often results in increased pollen production. Research on allergenic plants, such as ragweed, has
demonstrated that higher CO2 levels cause them to produce more pollen (Choi et al., 2021). These
environmental changes have health consequences. Pollen and mold exposure contribute to allergic rhinitis
and asthma (Sapkota et al., 2019), with a limited number of studies identifying relationships with
hospitalizations and mortality in people with underlying COPD (Idrose et al., 2022). A meta-analysis
found that each exposure increase of 10 grass pollen grains per m³ was associated with a 1.88% increase
in asthma emergency department visits (95% CI 0.94-2.82%) (Erbas et al., 2018), while high grass and
ragweed pollen concentrations were associated with chronic respiratory mortality in a Michigan study
(Larson et al., 2025). Elevated fungal spores have also been linked to increased asthma medication use,
symptom severity, and hospitalizations (D’Amato et al., 2020).
Prepublication Copy
6
Impacts on Public Welfare
Climate change influences public welfare in a multitude of ways that affect the places where we
live, work, and recreate; the food we consume; the water we drink and rely on to support agriculture and
energy production; and the air we breathe. This chapter focuses on climate change effects only on major
environmental systems that affect public welfare and are addressed in EPA (2009), including agriculture,
forests, grasslands, freshwaters, coastal oceans, and the built environment (energy, infrastructure, and
settlements). Aspects of climate change, notably temperature and precipitation, will manifest differently
in different regions of the United States. Thus, an evaluation of the evidence of climate change impacts on
public welfare is most appropriate when it considers the regional, if not subregional, level.
This chapter does not cover all ecosystems or public welfare impacts. Choices on what to
highlight were guided by the committee’s overall focus on impacts with more direct attribution to climate
conditions and more direct impacts on human health and well-being, as well as responses to a public
Request for Information. Nonetheless, the committee recognizes growing bodies of literature address
many other areas relevant to well-being, such as recreation, sport, hunting/fishing, and cultural heritage.
Likewise, there is a growing literature related to climate change and economics (see Box 6.1). Some
Prepublication Copy 57
examples of economic impacts of climate change are included, although an exhaustive review of
economic impacts was beyond the scope of this report.
Many of the effects on human health and welfare discussed in this report have associated economic impacts.
A growing number of studies since 2009 have estimated the economic effects associated with climate impacts on a
range of economic sectors. For example, a study that calculated the costs associated with lost agricultural
productivity is described in this chapter. For individuals who have directly experienced a climate impact, these
economic impacts may directly affect their earning potential, the value of property they own, or other factors that
contribute to their financial stability. Significant progress has been made since 2009 in analyzing sector-specific
and economy-wide impacts of climate change, though challenges remain in considering climate impacts in the
context of other economic drivers, such as changes in demographics, technology, and policy (NASEM, 2024).
Empirical evidence of economic impacts has been used to establish exposure-response functions, which relate
sector-specific economic impacts to climate indicators. These exposure-response functions can then be used to
estimate potential climate-related damages associated with different future climate scenarios. Studies in the
literature using empirical data to establish exposure-response relationships have expanded greatly since 2009 (e.g.,
Carleton and Greenstone, 2022; Clarke et al., 2018; Cromar et al., 2022; Depsky et al., 2023; Diaz, 2016; Moore
et al., 2017; Rode et al., 2021; Shindell et al., 2020). It is now possible to estimate future economy-wide damages
by aggregating these empirical exposure-response relationships for individual sectors (EPA, 2023). This is a
significant advance since 2009, when most economy-wide analyses used relationships between large-scale
economic indicators (e.g., Gross Domestic Product) and average changes in climate indicators (NASEM, 2017a).
An alternative approach that is also substantially advanced since 2009 uses improved empirical data to evaluate
economy-wide damages associated with each unit of change in the mean or variability of surface temperature or
precipitation (e.g., Kalkuhl et al., 2020; Kotz et al., 2021, 2022; Waidelich et al., 2024). Both the aggregated and
economy-wide approaches estimate significant costs associated with future climate change and, although they
have large uncertainties, their ranges overlap, providing increased confidence in their results.
Like the approach used in EPA (2009), this chapter includes discussion of direct impacts of
climate change but also considers some other key indirect effects on public welfare, to put climate change
impacts in the broader context of observed changes in environmental systems. The most significant and
well-documented climate change effects on public welfare and the changes in the evidence that have
occurred since 2009 are discussed here, including a better understanding of regional variability. In
addition to addressing public welfare topics broadly, a few cross-cutting topics where linkages to public
health (discussed Chapter 5) are strong are highlighted, including wildfire (discussed in Chapters 3 and
5), the nutritional status of food, harmful algal blooms, and toxic-laden sediments from exposed lake beds
carried as dust.
Ecosystems are complex, encompassing interactions among many biological communities with
important linkages with the physical environment and public welfare (IPBES, 2019). Climate is a key
controller of the structure and function of ecosystems. Climate change-driven shifts, particularly in
temperature and precipitation, are affecting the range of services that ecosystems and the built
environment provide. Linkages among temperature, precipitation, and other climate factors are explored
throughout this chapter.
Generally, across the public welfare areas discussed in EPA (2009), recent evidence has
strengthened the 2009 conclusions. New evidence has also led to improved understanding of the complex
interactions among climate and non-climate drivers that influence observed changes in ecosystems and
the built environment, and public welfare they support. In particular, the understanding of the regional
Prepublication Copy
variability of impacts and the complexity of other factors (e.g., land use, air quality, pests, and pathogens)
that interact with climate impacts has grown. Discussion of these interactions and variability are
addressed in this chapter.
Methane (CH4) emissions lead to increased ground-level ozone (see Chapter 3), which damages
many crops and trees. For the United States, analysis by the United Nations Environment Programme
provides estimates of crop yield losses driven by CH4 emissions (via induced climate, ozone, and CO2
changes), finding yield losses of roughly 1,750,000 tonnes of maize (corn), 340,000 tonnes of wheat,
60,000 tonnes of rice, and 790,000 tonnes of soybeans for every 100 million tonnes emitted (UNEP and
CCAC, 2021). 1 For context, 100 million tonnes represents about 25% of current anthropogenic CH4
emissions, and these losses represent 0.6-1.3% of global yields of these crops with values up to 3-4% for
individual countries (UNEP and CCAC, 2021). Non-CO2 GHGs such as nitrous oxide or fluorinated
gases, affect plants only via climate change.
Non-Climate Drivers
Important drivers of environmental system changes beyond those linked to climate can amplify or
mitigate public welfare effects. These non-climate drivers include land use and land cover change, pest
and pathogens, nitrogen deposition, and changes in air quality driven by air pollution emissions not
mediated through climate change. These non-climate drivers occur coincidently with climate change but
are highly variable in space and time. At the same time, some of these non-climate drivers are affected by
climate change (e.g., ozone and nitrogen deposition, noted in Chapter 3) and have also been shown to
either amplify or mitigate ecosystem response to a changing climate (e.g., Baron et al., 2013; Bytnerowicz
et al., 2007). Non-climate drivers contributing to the impacts discussed are noted throughout relevant
sections in this chapter.
Agriculture production and climate are intrinsically linked, and evidence collected since 2009
strengthens messages conveyed in EPA (2009). In this section, the committee details major effects of
climate and interacting non-climate drivers on crop production, agricultural pests and weeds, and
livestock.
1
These values were obtained from Tables 3.6a-d in UNEP and CCAC, 2021, and scaled to report values per 100
million tonnes.
Prepublication Copy
Increases in temperatures and variability in precipitation amount and intensity have negatively
affected agricultural production in the United States (Eck et al., 2020; Hatfield et al., 2011; Lesk et al.,
2022), although the extent of this impact varies by region. For the period of 1991-2017, temperature
related crop losses have resulted in $27 billion in crop insurance claims (Diffenbaugh et al., 2021).
Increased temperature has lengthened the frost-free days by 2 weeks since 1970, 2 however a longer
growing season increases the need for water and nutrients (fertilizer) to take advantage of the additional
crop growth potential.
Winters are also warming, as documented by the U.S. Department of Agriculture Plant Hardiness
Zone Map, 3 which provides guidance on where specific plants are most likely to thrive and has been
adjusted northward, most recently in both 2012 and 2023. These maps are developed using 30-year
averages of the lowest annual winter temperature at given locations, reflecting a trend in temperatures
with direct implications for plant growth.
Extreme heat, drought, and moisture excess are increasingly co-occurring within a single growing
season since 2000, resulting in up to 30% yield losses globally, with the United States noted as a region of
greatest losses (Lesk et al., 2022). Similarly, extreme heat events and warm nights have decreased yields,
and episodic temperature increases that exceed plant physiologic thresholds reduce yield and cause plant
stress throughout the life of the crop, especially during flowering (Hatfield et al., 2011; Schlenker and
Roberts, 2009).
Variability in temperature and precipitation effects has been observed across U.S. regions. For
example, excess heat and precipitation extremes within the growing season have negatively affected crop
yield in the southeast United States (Eck et al., 2020). The western United States has become hotter and
drier in recent years (1976-2019) with associated negative impacts on agriculture production and other
ecosystems (Su et al., 2021; Zhang et al., 2021).
Increasing temperatures across the United States have increased evaporation and plant
transpiration (water evaporation from the plant surface). This change leads to an increase in water deficits
and crop economic losses (Hatfield et al., 2011). In some parts of the United States, water will become
less available from both reductions in rainfall and increasing drawdown of water for irrigation (e.g., the
Ogallala region in the Great Plains). Hot-dry-windy events have significantly increased in the U.S. Great
Plains from 1982 to 2020. These events have resulted in a 4% yield reduction per 10 h of hot-dry-windy
conditions during the reproductive stage of wheat (Zhao et al., 2022).
In addition to direct effects of temperature and precipitation, changes in the nutritional value of
crops have been observed when grown under elevated CO2 conditions (see also Chapter 5 discussion).
Non-legume crop species often have lower protein content when grown under elevated CO2 (Kimball,
2010). C3 (cool season) grains and legumes have lower concentrations of zinc and iron when grown under
elevated CO2 while C4 (warm season) crops are less affected (Dietterich et al., 2015; Myers et al., 2014).
These nutritional changes affect dietary needs for both human food crops and livestock forage.
While many climate impacts have been observed, negative impacts might be more severe in the
absence of efforts to adapt or improve agricultural practices in response to observed change. These
adaptive measures include actions such as plant breeding, crop switching, soil management, and improved
technologies (e.g., irrigation, water conservation, precision agriculture).
Temperature is the single most important factor affecting insect ecology, epidemiology, the
number of generations per growing season, and insect distribution (Skendžić et al., 2021). Warmer
2
See https://www.epa.gov/climate-indicators/climate-change-indicators-length-growing-season (accessed
September 2, 2025).
3
See https://planthardiness.ars.usda.gov (accessed September 8, 2025).
Prepublication Copy
winters affect crops and weeds and also expand the potential habitable range of some insect and disease
pests. Plant pathogens are highly responsive to humidity and rainfall, as well as temperature (Lahlali et
al., 2024). Since 1960, data have documented a northward shift in pests (Bebber et al., 2013), and pests
are expected to reduce crop yield as a result of warming (Deutsch, 2018). Increased pests may also result
in more chemical applications at a cost to the farmer and the environment.
Many C3 weed species show substantial growth increases and resistance to herbicides when
grown at elevated CO2 levels (Ziska, 2003; Ziska et al., 1999). As a result, rising atmospheric CO2 could
lead to yield reductions when weed control is insufficient, potentially increasing the need for chemical
applications, increasing costs to farmers and chemicals in the environment.
The EPA (2009) discussion of livestock production is supported with new evidence and
strengthened by recent research findings. A growing body of evidence indicates that summer heat stress
has negative impacts on animal behavior. Livestock performance depends on their environment. The
direct effects of temperature, variable precipitation, and extreme events impact thermoregulation,
metabolism, and immune system function (Cheng et al., 2022). Heat stress increases susceptibility of
livestock to diseases and death and decreases weight gain, milk production, and reproduction rate. For
instance, from 2012-2016, milk yield was shown to be reduced by 1% due to heat stress resulting in $253
million in lost revenue across 19 states (Hutchins et al., 2025). Another study found that in 2010 lower
milk production due to heat stress resulted in up to $1.2 billion in losses to the dairy sector (Key et al.,
2014).
In addition to hotter summer temperatures generally, animals are experiencing more extreme heat
events and temperature swings which also impact animal behavior and stress. This heat exposure is more
acute for cattle on grazing lands because fewer options exist for mitigating heat effects. In addition, cattle
on grazing lands that are susceptible to fires bring the animals into close proximity to both fire and smoke.
Indirect effects of climate change on livestock relate to feed production (declines in and reduced
nutritional value), changes in water availability linked to shifting precipitation patterns, and increased
exposure to pests and parasites. In grazing systems, livestock production is reduced by lower forage
quality due to higher temperature, elevated CO2, and drought stress (Polley et al., 2013).
Climate change resulting from GHG emissions has impacted commercial marine fisheries in
every coastal region of the United States. Impacts on fisheries include losses in the abundance and quality
of harvested species and fisheries-related revenue and job loss (Fisher et al., 2021; Free et al., 2019;
Pershing et al., 2018). Changes in climate are not the only driver affecting fish populations but are an
additional stressor that can exacerbate other negative impacts and overwhelm and outpace even gold-
standard fisheries management regimes, such as North Pacific Fishery Management Council and National
Oceanic and Atmospheric Administration Fisheries in the Gulf of Alaska. Although some marine species
have benefitted from ocean warming, for example the northern stock of American lobster (Le Bris et al.,
2018), a variety of climate conditions related to warming have produced declines in other species. For
example, low sea ice conditions and a long period of warming temperatures in the Bering Sea led to
declines in stocks of Pacific cod and snow crab (Fedewa et al., 2020; Spies et al., 2020).
Negative responses have been observed in marine fisheries populations near the warmer edge of
their range, though a history of overfishing and other ocean changes are also contributing factors (Free et
al., 2019). Populations showing a positive response were those on the cold edge of the species range.
USGCRP (2023) noted that the incidence of disaster declarations for commercial fisheries rose from 1994
to 2019, and the majority of those disasters (more than 84%) were linked to extreme environmental
events.
Prepublication Copy
6.5 FORESTS
Forests are a critical resource for the United States, covering approximately a third of the nation’s
land area. The forest products industry represents about 4.7% of total U.S. manufacturing gross domestic
product and serves as an important manufacturing sector in the United States (AF&PA, 2022). Forest
cover has decreased slightly in the contiguous United States over the last 20 years largely due to
expansion of croplands and urbanization, which includes increases in development at the wildland-urban
interface (USGCRP, 2023). Forests are dominated by trees and woody vegetation and are commonly
situated in the headwaters of freshwater ecosystems (i.e., wetlands, rivers, streams, lakes). As a result,
forest and freshwater ecosystems are often intimately connected, with the structure and function of each is
dependent on the processes and resources the other supplies.
Forests provide a suite of services including marketable forest products, cleansing the atmosphere
of pollutants, retaining nutrients, influencing water supply, and flood and erosion control (USGCRP,
2023). Spending time in forests has also been shown to have positive health effects for people (Jimenez et
al., 2021). Additionally, forests also serve to regulate climate by the net removal of CO2 from the
atmosphere through photosynthesis and storage in tree biomass and soils. Forests also provide rich
biodiversity, aesthetics, recreation and cultural experiences (USGCRP, 2023).
Prepublication Copy
associations with trends in mean annual precipitation. Average annual growth and decadal survival
generally decreased with wetter conditions in the East and drier conditions in the West. Only eight species
considered were tolerant of increases in temperature. In the East, 24 species were found to be tolerant of
increases in precipitation and only seven in the West were tolerant of decreases in precipitation. There
were at least a few species that had a similar response (either positive or negative) across the contiguous
United States for growth and survival metrics.
The Forest Inventory and Analysis database has also recently been used to evaluate ozone
impacts on forest growth and survival of 88 tree species in the contiguous United States (Pavlovic et al.,
2025). As a whole, ozone exposure was generally below critical levels to impair tree growth, but
exceeded levels needed to protect survival for some species.
As discussed in Chapter 3 of this report, wildland fires are a growing climate concern for forested
ecosystems. At present the public health risks associated with these fires (discussed in Chapter 5 of this
report) are a dominant impact on public welfare, though other effects such as increased homeowners’
insurance premiums in fire-prone areas are growing.
As with crops, pests and pathogens are an important driver of tree growth and mortality and
fungal composition and function, whose impacts have been markedly altered and intensified by a
changing climate (Simler-Williamson et al., 2019). These impacts occur through tree physiology,
mortality and morbidity (Andrus et al., 2025; Cobb and Metz et al., 2017; Preston et al., 2016). Increases
in pathogens and insect pests cause changes in forest composition (Metz et al., 2012), disrupt food webs
(Ellison et al., 2005), and alter biogeochemical processes (Preston et al., 2016). Climate change alters the
survival rates of pests and pathogens (Simler-Williamson et al., 2019). Often rates of over-winter survival
limit outbreaks, but increasing winter temperatures have been linked to increasing pest occurrence and
impacts (McAvoy et al., 2017). Moreover, changes in temperature, humidity, and precipitation affect the
reproduction and growth rates of pests and pathogens. In addition to effects on pests and pathogens,
climate change can also impact the susceptibility of a tree host to infection, invasion, or damage resulting
in changes in physiology, morphology, and population or community structure.
6.6 GRASSLANDS
Grasslands account for approximately 29% of U.S. land area and serve as an important ecosystem
for livestock grazing and supporting wildlife. Grasslands are an important ecosystem for carbon storage;
globally, more than 30% of terrestrial carbon occurs in grasslands soils (Bai et al., 2022). Grasslands in
the Great Plains were estimated to contain 34.9% of the total carbon stocks in the region from 2001-2005
(Pendall et al., 2018). It has been estimated that U.S. grazing lands contribute 14.7% of the U.S. soil
carbon sequestration potential (Lal et al., 2003). Thus, grasslands are important for the economy and
ecosystem services.
Increasing temperature can reduce grass productivity in tallgrass prairie (Koerner et al., 2023).
Since 1984, the amount of plant biomass (known as annual net primary productivity) has increased in
Great Plains grasslands due to increased growing season precipitation (Reeves et al., 2021). However,
there are regional differences. The northern Great Plains may benefit from a longer growing season while
the southwestern Great Plains will likely show a decline due to increased drought, higher temperature,
and greater variability (McCollum et al., 2017). Since 2000, below average precipitation and above
average temperature have been observed in the southwestern United States, indicative of a changing
climate (Williams et al., 2023). This has reduced grassland productivity, with implications for grazing
livestock production. Rangeland grazing capacity in New Mexico has declined by 43% over a 52-year
period (1967-2018) due to higher growing season temperatures and increased frequency of drought
(McIntosh et al., 2019).
Prepublication Copy
Fire is a part of grassland ecology and historically has helped to suppress woody plants. However,
a combination of elevated CO2, increasing temperatures, and land and fire management practices is
contributing to the expansion of trees into grasslands (Morford et al., 2022) resulting in a loss of livestock
productivity.
Coastal ocean ecosystems are experiencing warmer waters, sea level rise, increasing pressures
from human development, and other stressors (May et al., 2023). These ecosystems play an important role
in protecting coastlines. Tidal wetlands, which include mangroves and salt marshes, provide crucial
habitats for fish and wildlife, and their dense vegetation helps to slow and absorb floodwaters, protecting
inland areas from storm surges and high tides. Wetlands serve as important nursery grounds and feeding
areas for many commercial fish species.
Sea level rise and increasing coastal hazards associated with climate change can drive tidal
wetland loss (Weis et al., 2021). Tidal wetlands can move landward to escape rising sea levels, a process
called inland migration. This can occur if there is space and time for the wetland to move inland before it
erodes or is submerged. The evidence of potential loss has grown since EPA (2009). A net loss of tidal
wetlands is expected throughout the United States, but the rate and extent of loss will vary significantly
from place to place depending on local conditions and inland migration. For example, loss and migration
of tidal wetlands linked to sea level rise has been observed in the Chesapeake Bay (Schieider et al., 2018),
Florida (Raabe and Stumpf, 2016), and New Jersey (Weis et al., 2021). Along the Pacific Coast, tidal
wetlands cannot migrate inland due to coastal development and steep topography, increasing the chances
of net tidal wetland loss due to sea level rise.
The amount and quality of water available for use by humans has direct impacts on welfare in a
variety of ways. Drought affects the production of food which leads to supply shortages and increased
prices. Increases in land area inundated by flood waters imply increases in losses of life and property.
Deteriorating water quality limits human use of water for multiple purposes, including drinking water and
recreation.
Impacts of climate change on water resources, including water quality and water availability,
droughts, and floods, are affected by regional hydroclimatology. For example, despite increasing air
temperatures everywhere in the United States, some regions, such as much of the East, are experiencing
significant increases in total precipitation, while in other regions, including parts of the West,
precipitation is decreasing. This section provides technical information about impacts and recent trends
and discusses the regional variability observed. Additional discussion of climate impacts on precipitation
and drought is provided in Chapter 3 of this report.
Water Quality
The quality of many streams and rivers across the contiguous United States is declining in
response to climate change, with implications for drinking water municipal use, energy, fisheries, and
other uses of freshwater. Water quality is affected by land cover and land use, and by many direct (e.g.,
sewage disposal) and indirect (e.g., use of fertilizers) human influences, creating complexity in
understanding climate and non-climate drivers of water quality change. A recent review of 965 case
studies indicated that 56% of observed water quality issues were related to climate change due to
increasing water temperatures and changes in low flow periods (van Vliet et al., 2023). Some substances,
such as nutrients and pharmaceuticals, show mostly increasing trends in concentrations, whereas others,
such as sediment, biochemical oxygen demand, and metals show a mixture of increasing and decreasing
trends in concentrations (van Vliet et al., 2023). Lakes have also been impacted by increasing
Prepublication Copy
temperatures in multiple ways, including a strengthening and lengthening period of thermal stratification,
enhanced depletion of dissolved oxygen in deeper waters, loss of habitat for cold-water fisheries, and
other threats such as spread of invasive species and loss of biodiversity (Jane et al., 2021, 2024; Woolway
et al., 2022).
A consequence of warming waters and enhanced stratification of lakes and coastal waters is
increases in harmful algal blooms (Lefebvre et al., 2025; Townhill et al., 2018; Trainer et al., 2020).
Additional changes in environmental conditions, such as increases in nutrients, can stimulate growth and
blooms of cyanobacteria, called harmful algal blooms (Chapra et al., 2017). Understanding and
documentation of climate-driven impacts of harmful algal blooms have increased greatly since 2009. In
freshwaters, cyanobacteria (also known as blue-green algae) occur naturally and are able to outcompete
other types of algae under warm water conditions (Cottingham et al., 2021). Some species of
cyanobacteria can release toxins when environmental conditions are favorable and the cyanobacteria
present can express genes that produce the toxins. These toxins can harm people and animals drinking or
recreating in contaminated waters or inhaling air near affected water sources (Plaas and Paerl, 2020).
Under extreme conditions, when drinking water sources are affected, closures or additional treatment of
the water supply may be needed.
In coastal waters, dinoflagellates or diatoms are the most common algae causing harmful algal
blooms (Anderson et al., 2021). In coastal waters, fish kills have been reported when water temperatures
are much higher than normal and associated with harmful algal bloom events. Shellfish contamination has
been documented, with toxins making shellfish unsafe to consume (e.g., Lefebvre et al., 2025; McCabe et
al., 2016).
For recreational fresh and coastal waters, many states post advisories in response to the
occurrence of harmful algal blooms, warning people against contact during recreational activities and the
potential for respiratory distress. Advisories have been listed for fresh and marine waters across the
United States, from Florida to Alaska. There has been a marked increase in the number of advisories
during the warm water months and in the annual total number of advisories since compilation was
initiated in 2015. 4 These increases likely reflect increases in harmful algal bloom events, increased
awareness of the problem, and improvements in monitoring efforts. Economic losses linked to harmful
algal bloom impacts on coastal fisheries and aquaculture in the United States have been estimated to be
tens of millions of dollars (Anderson et al., 2021; Jin et al., 2020).
Impacts of climate change on water supply and availability are complicated by the impacts of
non-climate factors, such as changes in water consumption. Overall water withdrawals for all uses in the
contiguous United States reported for 2015 were the lowest since 1970 (Warziniack et al., 2022). The
primary consumptive uses of water are for irrigation and thermoelectricity generation. Water withdrawals
for irrigation have remained relatively constant for decades despite increases in acreage under irrigation,
presumably because of increases in efficiency (Warziniack et al., 2022). Withdrawals for thermoelectric
power plant cooling have declined, again because of technological improvements (Warziniack et al.,
2022).
Hydroclimatic changes driven by atmospheric warming include changes in precipitation and
evapotranspiration demand. These changes interact with the land surface through exchanges of energy
and water. The consequences of these changes are reflected in spatial and temporal patterns of
precipitation, evapotranspiration, and soil moisture (Herrera et al., 2023). Changes of hydroclimate across
4
See the EPA Tracking CyanoHAB website at
https://storymaps.arcgis.com/stories/d4a87e6cdfd44d6ea7b97477969cb1dd (accessed September 3, 2025).
Prepublication Copy
the contiguous United States have changed the size and timing of rainfall over the past several decades
(Marvel et al., 2021).
Baseflow Drought
A meteorological drought is a period of low precipitation which has cascading effects. The lack
of precipitation delivered to the soil surface is accompanied by increased evaporation demand, which
leads to soil moisture drought. Additionally, reduced recharge to shallow groundwater leads to baseflow
(i.e., streamflow and groundwater flow) drought. The discussion in this section refers to baseflow drought
(see Chapter 3 for discussion of meteorological drought). EPA (2009) did not report significant evidence
of increases in drought due to climate change in the decades leading up to 2009. There is now some
evidence for increasing drought severity in some areas of the United States.
Historical data show that drought magnitude is increasing in some regions and decreasing in
others. For the period 1981-2020, drought duration and deficit was studied using reference (i.e., largely
unimpacted by significant land-use changes) watersheds and were found to have decreased in the north
and east and increased in the Southwest and south-central United States (Hammond et al., 2022). Alonso-
Vicario et al. (2025) included catchments that were impacted by agriculture and urbanization, but the
broad pattern for droughts across the United States was similar.
The impact of climate change in many regions may be confounded by the impact of other human
activities (Vicente-Serrano et al., 2022). Attribution of increasing drought trends to climate change is
difficult because the areas where these trends are observed overlap with areas of increasing water demand
and land cover change. In the Southwest, decreasing flows in the Colorado River have been attributed to
increased evapotranspiration driven by climate warming (Milly and Dunne, 2020). The observed 9.3%
flow decrease per degree Celsius of warming is likely to continue in the future although perhaps be
partially offset by increases in precipitation (Milly and Dunne, 2020).
Lag times between precipitation decreases and baseflow decreases are typically months and the
duration of baseflow droughts from months to years. Baseflow droughts in watersheds unimpacted by
human water use increased over the period from 1982 to 2012 in the mild temperate zone (Lee and Ajami,
2023). This zone consists generally of the Southeast, the eastern portion of the southern Great Plains, and
the West Coast. In watersheds impacted by groundwater withdrawals, links to climate change are indirect
(e.g., through the increased use of groundwater for irrigation). Meteorological droughts are linked to
groundwater level declines in broad areas of the United States. (Singh et al., 2025).
Lake levels have mostly increased across Alaska and the northern tier of the contiguous United
States and have decreased in the intermountain West and the Southeast over the past two decades (Feng et
al., 2022). Decreased water levels in terminal lakes in the Great Basin have been partially due to climate
change (Hall et al., 2023). Declines in water levels in these terminal lakes have direct implications for
human welfare as exposure of portions of the lake beds results in wind-blown sediments that contain toxic
metals (NASEM, 2020, see Chapter 5 for discussion of health effects of these toxins).
Floods
Across North America, the magnitude of extreme precipitation at the continental scale and at
broad regional scales has increased (Kirchmeier-Young and Zhang, 2020, see also Chapter 3 of this
report). Like other aspects of water resources, many factors influence flooding beyond climate. River
floods are affected by characteristics of the land and by both the amount and timing of precipitation.
Because of this combination of factors, at a regional scale, flood-hazard changes are not necessarily
directly linked to precipitation changes (Blöschl, 2022) and some report no clear signal of increases in
riverine floods despite the increases in extreme precipitation (Kundzewicz and Pinskwar, 2022). A more
nuanced analysis that separates areas of rain-induced flooding and areas of snowmelt-induced flooding
indicates that the annual maximum flood is increasing across the former and decreasing across the latter
(Zhang et al., 2022).
Prepublication Copy
In the western United States, data from reference watersheds indicate a decrease in the magnitude
and frequency of rain-on-snow flooding, an increase for convective storm flooding, and little change in
floods caused by other mechanisms (Huang et al., 2022). Patterns of flood magnitude trends studied
across the contiguous United States show land use and hydroclimate change to be equally important in
determining the trends (Kemter et al., 2023). Seasonal changes in flood frequency are also quite
heterogeneous spatially, but in general show more declining trends in spring and summer and more
increasing trends in autumn and winter (Gu et al., 2025). Trends in baseflow in rivers have been linked to
trends in annual floods in North America (Berghuijs and Slater, 2023). The complex interactions among
climate and land variables that result in river flows indicate that attribution of observed changes in flood
magnitudes and frequencies will be uncertain (Scussolini et al., 2024).
Annual maximum snowpack decreased significantly in the contiguous United States from 1982-
2016, and the snow season shortened by about a month (Zeng et al., 2018). These changes in snowpack
have important implications for soil moisture limitations during summer, wildfires (see Chapter 3) and
water supplies.
USGCRP (2023) highlights that flood hazards have disproportionate impacts on communities
across the country. Coastal communities, communities situated on rivers, agricultural, and fishing
communities experience more flood hazards (Edmonds et al., 2020; Thiault et al., 2019).
This section provides examples of climate change impacts on the built environment and describes
their links to public welfare. For energy, this includes discussion of energy production as well as changes
in demand associated with climate change. Similar to other welfare impacts discussed, impacts can be
wide-ranging and have considerable geographic variability. Where people live (e.g., coastal regions or in
or near forested areas) affects their vulnerability to some climate change impacts, such as coastal erosion
or wildfires, with attendant economic loss (Deilami et al., 2018).
Climate warming is increasing the number of cooling degree days and reducing heating degree
days 5 in the United States (EPA, 2024b). The result is significantly increased demand for air conditioning.
Air conditioning can help people remain in a temperature range that is comfortable and safe (see Box
5.1). Heat waves can result in both increased deaths and illness, with elderly and low-income populations
particularly vulnerable to increased heat in urban areas 6 (Qian and Liu, 2025). Urban areas can be
especially vulnerable to heat waves. Where the albedo (proportion of incident light that is reflected versus
absorbed by a surface) is low, incoming heat is absorbed by the built environment, creating Urban Heat
Islands, but attribution to increased GHG is difficult to establish (Martilli et al., 2000). Chapter 5 of this
report provides a more detailed discussion of temperature effects on public health and associated impacts.
Energy Production
Demand for electricity is continuing to increase as a result of population growth, higher incomes,
electrification of transportation, and internet services and data centers. Cooling needs due to warmer
5
Degree days are calculated using the assumption that when it is 65°F, neither heating or cooling is needed, then
calculating the difference between the daily temperature mean and 65°F. If the temperature mean is above 65°F, then
65 is subtracted from the mean to calculate Cooling Degree Days. If the temperature mean is below 65°F, the mean is
subtracted from 65 to calculate the Heating Degree Days. See https://www.weather.gov/key/climate_heat_cool
(accessed September 3, 2025).
6
See https://www.cdc.gov/heat-health/risk-factors/heat-and-older-adults-aged-65.html (accessed September
3, 2025).
Prepublication Copy
temperatures also has a significant impact on electricity demand. The peak demand for U.S. electricity
was set in late afternoon in July 2025 at 759,180 megawatts, representing a 2% increase from the previous
summer peak in 2024 (EIA, 2025). Increasing peaks require investment in new generating, transmission,
and distribution capacity.
Climate change is also increasing the costs of generating power (Bartos and Chester, 2015). As
noted in EPA (2009), warmer waters make the cooling of power plants less efficient, thus more costly,
and warmer temperatures make energy transmission less efficient. Extreme weather events coupled with
sea level rise in coastal areas create energy supply disruptions which require resiliency investments.
Hydroelectric power generation is susceptible to decreases due to droughts and decreased snowpacks.
Climate change is affecting numerous settlements, especially in coastal regions. 7 For example, in
Alaska, observed increases in coastal and riverbank erosion are causing damage and growing risks to
settlements in numerous communities (Huntington et al., 2023). Rapidly warming temperatures in the
Arctic region are leading to increased erosion through interactions with permafrost thaw, sea level rise,
declines in sea ice extent, the lengthening of open water period, and increased impacts of storms along
coastlines (Gibbs et al., 2021; Irrgang et al., 2022; Jones et al., 2020). In extreme cases, abandonment or
relocation of affected settlements is needed. Inland areas are also subject to permafrost thaw which can
cause ground subsidence and landslides, affecting settlements and infrastructure (Makopoulou et al.,
2025). The Gulf Coast is also experiencing similar settlement stresses due to sea level rise, hurricanes,
and erosion that can cause relocations of communities (NASEM, 2024a).
Nearly 40% of the U.S. population currently lives in coastal counties. Sea level rise (discussed in
Chapter 3) coupled with extreme weather has direct impacts on coastal infrastructure and settlements,
including flooding, erosion, wind damage, and saltwater incursion to water supplies and cropland. For
example, saltwater incursion and the pressure of the water even when it is not moving may damage
structural foundations (Abdelhafez et al., 2022).
Increased wildfire severity (discussed in Chapter 3) is impacting infrastructure and settlements. In
the western United States, structural losses due to wildfires increased over 200% from the decade 1999-
2000 to 2000-2010 (Higuera et al., 2023). Wildland-urban interfaces are growing as settlements expand
spatially into fire-prone areas, posing significant risks to structures and human health (NASEM, 2022; see
Chapter 5). The cost of insurance is also rising while the ability to retain and obtain insurance is declining
in areas with wildfire risks (Auer, 2024).
Climate change also affects U.S. highway infrastructure, where impacts are driven by
temperature, precipitation, sea level rise, and hurricanes (see discussion of hurricanes in Chapter 3). Table
6.1 details highway impacts (TRB, 2014). Increased costs are already being observed from effects such as
softening asphalt due to high temperatures (Sias et al., 2025). Thermal expansion on roadways and bridges
is also a significant issue (Zhiyuan et al., 2025). Other infrastructure systems experience similar impacts,
including energy, public transportation, railroads, ports, military facilities, and water infrastructure.
7
See https://oceanservice.noaa.gov/facts/population.html (accessed September 3, 2025).
Prepublication Copy
• Bridges subject to extra stresses • Vehicle overheating and increased risk of tire blowouts.
through thermal expansion and • Rising transportation costs (increase need for
increased movement. refrigeration).
• Materials and load restrictions can limit transportation
operations.
• Closure of roads because of increased wildfires.
Change in range of • Shorter snow and ice season. • Decrease in frozen precipitation would improve mobility
maximum and • Reduced frost heave and road and safety of travel through reduced winter hazards, reduce
minimum temperature damage. Later freeze and earlier snow and ice removal costs, decrease need for winter road
thaw of structures because of maintenance, and result in less pollution from road salt, and
shorter freeze-season lengths decrease corrosion of infrastructure and vehicles.
Increased freeze–thaw • Longer road construction season in colder locations.
conditions in selected locations • Vehicle load restrictions in place on roads to minimize
creating frost heaves and structural damage due to subsidence and the loss of bearing
potholes on road and bridge capacity during spring thaw period (restrictions likely to
surfaces. expand in areas with shorter winters but longer thaw
• Increased slope instability, seasons).
landslides, and shoreline • Roadways built on permafrost likely to be damaged due to
erosion from permafrost lateral spreading and settlement of road embankments.
thawing leads to damaging • Shorter season for ice roads.
roads and bridges due to
foundation settlement (bridges
and large culverts are
particularly sensitive to
movement caused by thawing
permafrost).
• Hotter summers in Alaska lead
to increased glacial melting and
longer periods of high stream
flows, causing both increased
sediment in rivers and scouring
of bridge supporting piers and
abutments.
Precipitation
Greater changes in • If more precipitation falls as • Regions with more precipitation could see increased
precipitation levels rain rather than snow in winter weather-related accidents, delays, and traffic disruptions
and spring, there will be an (loss of life and property, increased safety risks, increased
increased risk of landslides, risks of hazardous cargo accidents).
slope failures, and floods from • Roadways and underground tunnels could close due to
the runoff, causing road flooding and mudslides in areas deforested by wildfires.
washouts and closures as well • Increased wildfires during droughts could threaten roads
as the need for road repair and directly or cause road closures due to fire threat or reduced
reconstruction. visibility.
• Increasing precipitation could • Clay subsurfaces for pavement could expand or contract in
lead to soil moisture levels prolonged precipitation or drought, causing pavement
becoming too high (structural heave or cracking.
integrity of roads, bridges, and
tunnels could be compromised
leading to accelerated
deterioration).
• Less rain available to dilute
surface salt may cause steel
reinforcing in concrete
structures to corrode.
• Road embankments could be at
risk of subsidence/heave.
• Subsurface soils may shrink
because of drought.
Prepublication Copy
Increased intense • Heavy winter rain with • The number of road closures due to flooding and washouts
precipitation, other change accompanying mudslides can will likely rise.
in storm intensity (except damage roads (washouts and • Erosion will occur at road construction project sites as
hurricanes) undercutting), which could lead heavy rain events take place more frequently.
to permanent road closures. • Road construction activities could be disrupted.
• Heavy precipitation and • Increases in weather-related highway accidents, delays, and
increased runoff can cause traffic disruptions are likely.
damage to tunnels, culverts, • Increases in landslides, closures or major disruptions of
roads in or near flood zones, roads, emergency evacuations, and travel delays are likely.
and coastal highways. • Increased wind speeds could result in loss of visibility from
• Bridges are more prone to drifting snow, loss of vehicle stability/maneuverability,
extreme wind events and lane obstruction (debris), and treatment chemical
scouring from higher stream dispersion.
runoff. • Lightning/electrical disturbance could
• Bridges, signs, overhead cables, disrupt transportation electronic infrastructure
and tall structures could be at and signaling, pose risk to personnel, and delay
risk from increased wind maintenance activity.
speeds.
Sea Level
Sea-level rise • Erosion of coastal road base and • Coastal road flooding and damage resulting from sea-level
undermining of bridge supports rise and storm surge.
due to higher sea levels and • Increased exposure to storm surges.
storm surges. Temporary and • More frequent and severe flooding of underground tunnels
permanent flooding of and other low-lying infrastructure.
roads and tunnels due to rising
sea levels.
• Encroachment of saltwater
leading to accelerated
degradation of tunnels (reduced
life expectancy, increased
maintenance costs and potential
for structural failure during
extreme events).
• Further coastal erosion due to
the loss of coastal wetlands and
barrier islands.
• removing natural protection
from wave action.
Hurricanes
Increased hurricane • Increased infrastructure damage • More frequent flooding of coastal roads.
intensity and failure (highway and bridge • More transportation interruptions (storm debris on roads
decks being displaced). can damage infrastructure and interrupt travel and
shipments of goods).
• More coastal evacuations.
Prepublication Copy
References
Abadi, A.M., Y. Gwon, M.O. Gribble, J.D. Berman, R. Bilotta, M. Hobbins, and J.E Bell. 2022. “Drought
and All-Cause Mortality in Nebraska from 1980 to 2014: Time-Series Analyses by Age, Sex, Race,
Urbanicity and Drought Severity.” Science of The Total Environment 840:156660.
https://doi.org/10.1016/j.scitotenv.2022.156660.
Abatzoglou, J.T., and A.P. Williams. 2016. “Impact of Anthropogenic Climate Change on Wildfire
Across Western US Forests.” Proceedings of the National Academy of Sciences 113(42):11770–
11775. https://doi.org/10.1073/pnas.1607171113.
Abbatt, J.P.D., and C. Wang. 2020. “The Atmospheric Chemistry of Indoor Environments.”
Environmental Science: Processes & Impacts 22:25–48. https://doi.org/10.1039/C9EM00386J
Abdelhafez, M.A., B. Ellingwood, and H. Mahmoud. 2022. “Hidden Costs to Building Foundations Due
to Sea Level Rise in a Changing Climate.” Scientific Reports 12(14020).
https://doi.org/10.1038/s41598-022-18467-3.
Abolhasani, E., V. Hachinski, N. Ghazaleh, M.R. Azarpazhooh, N. Mokhber, and J. Martin. 2023. “Air
Pollution and Incidence of Dementia: A Systematic Review and Meta-Analysis.” Neurology
100(2):e242–e254. https://doi.org/10.1212/WNL.0000000000201419.
Abram, N.J., A. Purich, M.H. England, F.S. McCormack, J.M. Strugnell, D.M. Bergstrom, T.R. Vance, et
al. 2025. “Emerging Evidence of Abrupt Changes in the Antarctic Environment.” Nature 644:621–
633. https://doi.org/10.1038/s41586-025-09349-5.
Achakulwisut, P., L.J. Mickley, and S.C. Anenberg. 2018. “Drought-Sensitivity of Fine Dust in the US
Southwest: Implications for Air Quality and Public Health under Future Climate Change.”
Environmental Research Letters 13(5):054025. https://doi.org/10.1088/1748-9326/aabf20.
Agache, I., I. Annesi‐Maesano, L. Cecchi, B. Biagioni, K.F. Chung, B. Clot, G. D’Amato, et al. 2024.
“EAACI Guidelines on Environmental Science for Allergy and Asthma: The Impact of Short‐Term
Exposure to Outdoor Air Pollutants on Asthma‐Related Outcomes and Recommendations for
Mitigation Measures.” Allergy 79(7):1656–1686. https://doi.org/10.1111/all.16103.
Aggarwal, S., J.K. Hu, J.A. Sullivan, R.M. Parks, and R.C. Nethery. 2025. “Severe Flooding and Cause-
Specific Hospitalisation Among Older Adults in the USA: A Retrospective Matched Cohort
Analysis.” The Lancet. Planetary Health 9(7):101268. https://doi.org/10.1016/S2542-5196(25)00132-
9.
Aguilera, R., T. Corringham, A. Gershunov, and T. Benmarhnia. 2021. “Wildfire Smoke Impacts
Respiratory Health More than Fine Particles from Other Sources: Observational Evidence from
Southern California.” Nature Communications 12(1):1493. https://doi.org/10.1038/s41467-021-
21708-0.
Ahn, J. 2023. “Chapter 2 of the Working Group I Contribution to the IPCC Sixth Assessment Report -
data for Figure 2.4 (v20221219).” NERC EDS Centre for Environmental Data Analysis.
https://dx.doi.org/10.5285/60eeb3cce51a457cb5ee1c577a0c8674
Alahmad, B., H. Khraishah, D. Royé, A.M. Vicedo-Cabrera, Y. Guo, S.I. Papatheodorou, S. Achilleos, et
al. 2023. “Associations Between Extreme Temperatures and Cardiovascular Cause-Specific
Mortality: Results From 27 Countries.” Circulation, 147(1):35–46.
https://doi.org/10.1161/CIRCULATIONAHA.122.061832
Alahmad, B., A. Tobias, P. Masselot, and A. Gasparrini. 2025. “Are There More Cold Deaths Than Heat
Deaths?” The Lancet: Planetary Health 9(3):e170–e171. https://doi.org/10.1016/S2542-
5196(25)00054-3.
Alari, A., J. Ballester, C. Milà, T. Benmarhnia, M. Sofiev, A. Uppstu, R. Hänninen, and C. Tonne. 2025.
“Quantifying the Short-Term Mortality Effects of Wildfire Smoke in Europe: A Multicountry
Prepublication Copy 71
Epidemiological Study in 654 Contiguous Regions.” The Lancet: Planetary Health 9(8):101296.
https://doi.org/10.1016/j.lanplh.2025.101296.
Almetwally, A.A., M. Bin-Jumah, and A.A. Allam. 2020. “Ambient Air Pollution and its Influences on
Human Health and Welfare: An Overview.” Environmental Science and Pollution Research
27(20):24815–24830. https://doi.org/10.1007/s11356-020-09042-2.
Alonso-Vicario, S., G.M. Hornberger, M. Mazzoleni, and M. Garcia. 2025. “Drivers and Trends of
Streamflow Droughts in Natural and Human-Impacted Basins Across the Contiguous United States.”
Journal of Hydrology 655:132908. https://doi.org/10.1016/j.jhydrol.2025.132908.
Amdur, T., and P. Huybers. 2025. “Negative Trend in Total Solar Irradiance over the Satellite Era.”
Proceedings of the National Academy of Sciences 122(11):e2417155122.
https://doi.org/10.1073/pnas.2417155122.
American Forest and Paper Association (AF&PA). 2022. “Our Economic Impact.” American Forest and
Paper Association. https://www.afandpa.org/statistics-resources/our-economic-impact.
Anand, S., A. Staniec, M. Montez-Rath, and P. Vlahos. 2020. “Using GIS Mapping to Track Hot Spots of
Kidney Disease in California.” New England Journal of Medicine 382(23):2265–2267.
https://doi.org/10.1056/NEJMc2001023.
Anderegg, W.R.L., J.T. Abatzoglou, L.D.L. Anderegg, L. Bielory, P.L. Kinney, and L. Ziska. 2021.
“Anthropogenic Climate Change is Worsening North American Pollen Seasons.” Proceedings of the
National Academy of Sciences 118(7):e2013284118. https://doi.org/10.1073/pnas.2013284118.
Anderson, D.M., E. Fensin, C.J. Gobler, A.E. Hoeglund, K.A. Hubbard, D.M. Kulis, J.H. Landsberg, et
al. 2021. “Marine Harmful Algal Blooms (HABs) in the United States: History, Current Status, and
Future Trends.” Harmful Algae 102. https://doi.org/10.1016/j.hal.2021.101975
Andrus, R.A., J.A. Hicke, and A.J.H. Meddens. 2025. “Spatiotemporal Characteristics of Tree Mortality
from Bark Beetle Outbreaks Vary Within and Among Bark Beetle-Host Tree Associations in the
Western United States.” Forest Ecology and Management 576:122382.
https://doi.org/10.1016/j.foreco.2024.122382.
Anenberg, S.C., S. Haines, E. Wang, N. Nassikas, and P.L. Kinney. 2020. “Synergistic Health Effects of
Air Pollution, Temperature, and Pollen Exposure: A Systematic Review of Epidemiological
Evidence.” Environmental Health 19(1):130. https://doi.org/10.1186/s12940-020-00681-z.
Anenberg, S.C., and C. Kalman. 2019. “Extreme Weather, Chemical Facilities, and Vulnerable
Communities in the U.S. Gulf Coast: A Disastrous Combination.” GeoHealth 3(5):122–126.
https://doi.org/10.1029/2019GH000197.
Apurv, T., and X. Cai. 2021. “Regional Drought Risk in the Contiguous United States.” Geophysical
Research Letters 48(5). https://doi.org/10.1029/2020GL092200.
Archer, E.J., C. Baker-Austin, T.J. Osborn, N.R. Jones, J. Martínez-Urtaza, J. Trinanes, J.D. Oliver,
F.J.C. González, and I.R. Lake. 2023. “Climate Warming and Increasing Vibrio vulnificus Infections
in North America.” Scientific Reports 13(1):3893. https://doi.org/10.1038/s41598-023-28247-2
Armstrong McKay, D.I., A. Staal, J.F. Abrams, R. Winkelmann, B. Sakshewski, S. Loriani, I. Fetzer, S.E.
Cornell, J. Rockström, and T.M. Lenton. 2022. “Exceeding 1.5°C Global Warming Could Trigger
Multiple Climate Tipping Points.” Science 377(6611). https://doi.org/10.1126/science.abn7950.
Arregui-García, B., C. Ascione, A. Pera, B. Wang, D. Stocco, C.J. Carlson, S. Bansal, E. Valdano, and G.
Pullano. 2025. “Disruption of Outdoor Activities Caused by Wildfire Smoke Shapes Circulation of
Respiratory Pathogens.” PLOS Climate 4(6):e0000542.
https://doi.org/10.1371/journal.pclm.0000542.
Arrhenius, S. 1896. “On the Influence of Carbonic Acid in the Air upon the Temperature of the Ground.”
Philosophical Magazine and Journal of Science 5(41):237–276.
https://www.rsc.org/images/arrhenius1896_tcm18-173546.pdf.
Auer, M.R. 2024. “Wildfire Risk and Insurance: Research Directions for Policy Scientists.” Policy
Sciences 57:459–484. https://doi.org/10.1007/s11077-024-09528-7.
Prepublication Copy
References 73
Baharav, Y., L. Nichols, A. Wahal, O. Gow, K. Shickman, M. Edwards, and K. Huffling. 2023. “The
Impact of Extreme Heat Exposure on Pregnant People and Neonates: A State of the Science Review.”
Journal of Midwifery & Women’s Health 68(3):324–332. https://doi.org/10.1111/jmwh.13502.
Bai, Y., and M.F. Cotrufo. 2022. “Grassland Soil Carbon Sequestration: Current Understanding,
Challenges, and Solutions.” Science 377:603-608. https://doi.org/10.1126/science.abo2380.
Balbus, J.M., J.B. Greenblatt, R. Chari, D. Millstein, and K.L. Ebi. 2015. “A Wedge-Based Approach to
Estimating Health Co-Benefits of Climate Change Mitigation Activities in the United States.”
Climatic Change 127:199–210. https://doi.org/10.1007/s10584-014-1262-5.
Baldwin, J.W., T. Benmarhnia, K.L. Ebi, O. Jay, N.J. Lutsko, and J.K. Vanos. 2023. “Humidity’s Role in
Heat-Related Health Outcomes: A Heated Debate.” Environmental Health Perspectives
131(5):055001. https://doi.org/10.1289/EHP11807.
Barbani, S.M. 2025. “The Indirect Consequences of Climate Change and Extreme Environmental Events
on Mental Health: A Literature Review.” Journal of Public Health 1–14.
https://doi.org/10.1007/s10389-025-02513-1.
Baron, J.S., E.K. Hall, B.T. Nolan, J.C. Finlay, E.S. Bernhardt, J.A. Harrison, F. Chan, and E.W. Boyer.
2013. “The Interactive Effects of Excess Reactive Nitrogen and Climate Change on Aquatic
Ecosystems and Water Resources of the United States.” Biogeochemistry 114(1–3):71–92.
https://doi.org/10.1007/s10533-012-9788-y.
Bartos, M., and M. Chester. 2015. “Impacts of Climate Change on Electric Power Supply in the Western
United States.” Nature Climate Change 5:748–752. https://doi.org/10.1038/nclimate2648.
Bates, N.R., and R.J. Johnson. 2020. “Acceleration of Ocean Warming, Salinification, Deoxygenation and
Acidification in the Surface Subtropical North Atlantic Ocean.” Communications Earth &
Environment 1:33. https://doi.org/10.1038/s43247-020-00030-5
Baum, A., M.L. Barnett, J. Wisnivesky, and M.D. Schwartz. 2019. “Association Between a Temporary
Reduction in Access to Health Care and Long-term Changes in Hypertension Control Among
Veterans After a Natural Disaster.” JAMA Network Open 2(11):e1915111.
https://doi.org/10.1001/jamanetworkopen.2019.15111.
Baxter, B.K., and J.K. Butler. 2020. “Climate Change and Great Salt Lake.” In Great Salt Lake Biology.
B. K. Baxter & J. K. Butler, eds. 23–52. Springer International Publishing.
https://doi.org/10.1007/978-3-030-40352-2_2.
Bebber, D.P., M.A.T. Ramotowski, and S.J. Gurr. 2013. “Crop Pests and Pathogens Move Polewards in a
Warming World.” Nature Climate Change 3(11):985–988. https://doi.org/10.1038/nclimate1990.
Beggs, P.J. 2024. “Thunderstorm Asthma and Climate Change.” JAMA 331(10):878.
https://doi.org/10.1001/jama.2023.26649.
Bell, M.L., D.L. Davis, L.A. Cifuentes. A.J. Krupnick, R.D. Morgenstern, and G.D. Thurston. 2008.
“Ancillary Human Health Benefits of Improved Air Quality Resulting from Climate Change
Mitigation.” Environmental Health 7:41. https://doi.org/10.1186/1476-069X-7-41.
Bellizzi, S., C. Popescu, C.M. Panu Napodano, M. Fiamma, and L. Cegolon. 2023. “Global Health,
Climate Change and Migration: The Need for Recognition of “Climate Refugees.” Journal of Global
Health 13:03011. https://doi.org/10.7189/jogh.13.03011.
Berghuijs, W.R., and L.J. Slater. 2023. “Groundwater Shapes North American River Floods.”
Environmental Research Letters 18(3):034043. https://doi.org/10.1088/1748-9326/acbecc.
Berman, J.D., M.R. Ramirez, J.E. Bell, R. Bilotta, F. Gerr, and N.B. Fethke. 2021. “The Association
Between Drought Conditions and Increased Occupational Psychosocial Stress among U.S. Farmers:
An Occupational Cohort Study.” Science of The Total Environment 798:149245.
https://doi.org/10.1016/j.scitotenv.2021.149245.
Best Rogowski, C.B., C. Bredell, Y. Shi, A. Tien-Smith, M. Szybka, K.W. Fung, L. Hong, et al. 2025.
“Long-Term Air Pollution Exposure and Incident Dementia: A Systematic Review and Meta-
Analysis.” The Lancet Planetary Health 9(7):101266. https://doi.org/10.1016/S2542-5196(25)00118-
4.
Prepublication Copy
Blöschl, G. 2022. “Three Hypotheses on Changing River Flood Hazards.” Hydrology and Earth System
Sciences 26(19):5015–5033. https://doi.org/10.5194/hess-26-5015-2022.
Blunden, J., and J. Reagan. 2025. “State of the Climate in 2024.” Bulletin of the American Meteorological
Society 106(8):Si–S513. https://doi.org/10.1175/2025BAMSStateoftheClimate.1.
Bragg-Gresham, J., H. Morgenstern, V. Shahinian, B. Robinson, K. Abbott, and R. Saran. 2020. “An
Analysis of Hot Spots of ESRD in the United States: Potential Presence of CKD of Unknown Origin
in the USA?” Clinical Nephrology 93(1):113–119. https://doi.org/10.5414/CNP92S120.
Browman, H.I. 2016. “Applying Organized Scepticism to Ocean Acidification Research.” ICES Journal
of Marine Science 73(3):529–536. https://doi.org/10.1093/icesjms/fsw010.
Brumfield, K.D., M. Usmani, D.M. Long, H.A. Lupari, R.K. Pope, A.S. Jutla, A. Huq, and R.R. Colwell.
2025. “Climate Change and Vibrio: Environmental Determinants for Predictive Risk Assessment.”
Proceedings of the National Academy of Sciences 122(33):e2420423122.
https://doi.org/10.1073/pnas.2420423122.
Buonocore J.J., K.F. Lambert, D. Burtraw, S. Sekar, and C.T. Driscoll. 2016. “An Analysis of Costs and
Health Co-Benefits for a U.S. Power Plant Carbon Standard.” PLoS ONE
11(6):e0156308.https://doi.org/10.1371/journal.pone.0156308.
Burbank, A.J. 2025. “Climate Change and the Future of Allergies and Asthma.” Current Allergy and
Asthma Reports 25(1):20. https://doi.org/10.1007/s11882-025-01201-0.
Burkart, K.G., M. Brauer, A.Y. Aravkin, W.W. Godwin, S.I. Hay, J. He, V.C. Iannucci, et al. 2021.
“Estimating the Cause-Specific Relative Risks of Non-Optimal Temperature on Daily Mortality: A
Two-Part Modelling Approach Applied to the Global Burden of Disease Study.” The Lancet
398(10301):685–697. https://doi.org/10.1016/S0140-6736(21)01700-1.
Byrne, B., J. Liu, K.W. Bowman, M. Pascolini-Campbell, A. Chatterjee, S. Pandey, K. Miyazaki, et al.
2024. “Carbon Emissions from the 2023 Canadian Wildfires.” Nature 633:835–839.
https://doi.org/10.1038/s41586-024-07878-z
Bytnerowicz, A., K. Omasa, and E. Paoletti. 2007. “Integrated Effects of Air Pollution and Climate
Change on Forests: A Northern Hemisphere Perspective.” Environmental Pollution 147(3):438–445.
https://doi.org/10.1016/j.envpol.2006.08.028.
Campbell, J.L., C.T. Driscoll, J.A. Jones, E.R. Boose, H.A. Dugan, P.M. Groffman, C.R. Jackson, et al.
2022. “Forest and Freshwater Ecosystem Responses to Climate Change and Variability at US LTER
Sites.” BioScience 72(9):851–870. https://doi.org/10.1093/biosci/biab124.
Canadell, J.G., P.M.S. Monteiro, M.H. Costa, L. Cotrim da Cunha, P.M. Cox, A.V. Eliseev, S. Henson, et
al. 2021. “Global Carbon and other Biogeochemical Cycles and Feedbacks.” In Climate Change
2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of
the Intergovernmental Panel on Climate Change (IPCC). Cambridge University Press, 673–816.
https://doi.org/10.1017/9781009157896.007.
Capron, E., S.O. Rasmussen, T.J. Popp, T. Erhardt, H. Fischer, A. Landais, J.B. Pedro, et al. 2021. “The
Anatomy of Past Abrupt Warmings Recorded in Greenland Ice.” Nature Communications 12:2106.
https://doi.org/10.1038/s41467-021-22241-w.
Carleton, T., and M. Greenstone. 2022. “A Guide to Updating the US Government’s Social Cost of
Carbon.” Review of Environmental Economics and Policy 16(2):196–218.
https://doi.org/10.1086/720988.
Carleton, T., A. Jina, M. Delgado, M. Greenstone, T. Houser, S. Hsiang, A. Hultgren, et al. 2022.
“Valuing the Global Mortality Consequences of Climate Change Accounting for Adaptation Costs
and Benefits.” The Quarterly Journal of Economics 137(4):2037–2105.
https://doi.org/10.1093/qje/qjac020.
Casas, M.C., G.A. Schmidt, R.L. Miller, C. Orbe, K. Tsigaridis, L.S. Nazarenko, S.E. Bauer, and D.T.
Shindell. 2023. “Understanding Model-Observation Discrepancies in Satellite Retrievals of
Atmospheric Temperature Using GISS ModelE.” Journal of Geophysical Research:
Atmospheres 128(1):e2022JD037523. https://doi.org/10.1029/2022JD037523.
Prepublication Copy
References 75
Casson, N.J., A.R. Contosta, E.A. Burakowski, J.L. Campbell, M.S. Crandall, I.F. Creed, M.C. Eimers, et
al. 2019. “Winter Weather Whiplash: Impacts of Meteorological Events Misaligned with Natural and
Human Systems in Seasonally Snow-Covered Regions.” Earth’s Future 7:1434–
1450. https://doi.org/10.1029/2019EF001224
Centers for Disease Control and Prevention (CDC). 2004. Lyme Disease - United States, 2001–2002.
MMWR 53(17):365–369. https://www.cdc.gov/mmwr/preview/mmwrhtml/mm5317a4.htm.
Chandanpurkar, H.A., J.S. Famiglietti, K. Gopalan, D.N. Wiese, Y. Wada, K. Kakinuma, J.T. Reager, and
F. Zhang. 2025. “Unprecedented Continental Drying, Shrinking Freshwater Availability, and
Increasing Land Contributions to Sea Level Rise.” Science Advances 11(30).
https://doi.org/10.1126/sciadv.adx0298.
Chang, K.-L., B.C. McDonald, C. Harkins, and O.R. Cooper. 2025. “Surface Ozone Trend Variability
Across the United States and the Impact of Heat Waves (1990–2023).” Atmospheric Chemistry and
Physics 25(10):5101–5132. https://doi.org/10.5194/acp-25-5101-2025.
Chapra, S.C., B. Boehlert, C. Fant, V.J. Bierman, Jr., J. Henderson, D. Mills, D.M.L. Mas, et al. 2017.
“Climate Change Impacts on Harmful Algal Blooms in U.S. Freshwaters: A Screening-Level
Assessment.” Environmental Science and Technology 51(16):8933–8943.
https://doi.org/10.1021/acs.est.7b01498
Chen, C., B. Su, Y. Wang, Y. Zhao, Y. Wu, J. Li, Y. Luo, and X. Zheng. 2025. “Cumulative Effect of
PM2.5 Chemical Components Surpasses PM2.5 Mass on Hypertension in Older Adults: A China-
Based National Analysis.” Journal of Environmental Sciences 157:782–792.
https://doi.org/10.1016/j.jes.2025.01.014.
Chen, C., L. Schwarz, N. Rosenthal, M.E. Marlier, and T. Benmarhnia. 2024. “Exploring Spatial
Heterogeneity in Synergistic Effects of Compound Climate Hazards: Extreme Heat and Wildfire
Smoke on Cardiorespiratory Hospitalizations in California.” Science Advances 10(5):eadj7264.
https://doi.org/10.1126/sciadv.adj7264.
Chen, K., R. Dubrow, S. Breitner-Busch, K. Wolf, J. Linseisen, T. Schmitz, M. Heier, et al. 2022.
“Triggering of Myocardial Infarction by Heat Exposure is Modified by Medication Intake.” Nature
Cardiovascular Research 1:727–731. https://doi.org/10.1038/s44161-022-00102-z.
Cheng, L., J.P. Abraham, Z. Hausfather, and K.E. Trenberth. 2019. “How Fast are the Oceans Warming?”
Science 363(6423):128–129. https://doi.org/10.1126/science.aav7619.
Cheng, M., B. McCarl, and C. Fei. 2022. “Climate Change and Livestock Production: A Literature
Review.” Atmosphere 13(1):140. https://doi.org/10.3390/atmos13010140
Chew, G.L., J. Wilson, F.A. Rabito, F. Grimsley, S. Iqbal, T. Reponen, M.L. Muilenberg, P.S. Thorne,
D.G. Dearborn, and R.L. Morley. 2006. “Mold and Endotoxin Levels in the Aftermath of Hurricane
Katrina: A Pilot Project of Homes in New Orleans Undergoing Renovation.” Environmental Health
Perspectives 114(12):1883–1889. https://doi.org/10.1289/ehp.9258.
Childs, M.L., J. Li, J. Wen, S. Heft-Neal, A. Driscoll, S. Wang, C.F. Gould, M. Qiu, J. Burney, and M.
Burke. 2022. “Daily Local-Level Estimates of Ambient Wildfire Smoke PM2.5 for the Contiguous
US.” Environmental Science & Technology 56(19):13607–13621.
https://doi.org/10.1021/acs.est.2c02934.
Choi, H.M., S. Heo, D. Foo, Y. Song, R. Stewart, J. Son, and M.L. Bell. 2024. “Temperature, Crime, and
Violence: A Systematic Review and Meta-Analysis.” Environmental Health Perspectives
132(10):106001. https://doi.org/10.1289/EHP14300.
Choi, Y.J., K.S. Lee, and J.W. Oh. 2021. “The Impact of Climate Change on Pollen Season and Allergic
Sensitization to Pollens.” Immunology and Allergy Clinics 41(1):97–109.
https://doi.org/10.1016/j.iac.2020.09.004.
Chou, C., J.D. Neelin, C.-A. Chen, and J.-Y. Tu. 2009. “Evaluating the ‘Rich-Get-Richer’ Mechanism in
Tropical Precipitation Change Under Global Warming.” Journal of Climate 22(8):1982–2005.
https://doi.org/10.1175/2008JCLI2471.1
Prepublication Copy
Chu, L., J.L. Warren, E.S. Spatz, S. Lowe, Y. Lu, X. Ma, J.S. Ross, H.M. Krumholz, and K. Chen. 2025.
“Floods and Cause-Specific Mortality in the United States Applying a Triply Robust Approach.”
Nature Communications 16(1):2853. https://doi.org/10.1038/s41467-025-58236-0.
Cissé, G., R.McLeman, H. Adams, P. Aldunce, K. Bowen, D. Campbell-Lendrum, S. Clayton, et al. 2022.
“Health, Wellbeing, and the Changing Structure of Communities.” In Climate Change 2023: Impacts,
Adaptation, and Vulnerability. Contribution of Working Group II to the Sixth Assessment Report of
the Intergovernmental Panel on Climate Change. Pörtner, H.-O., D.C. Roberts, M. Tignor, E.S.
Poloczanska, K. Mintenbeck, A. Alegría, M. Craig, S. Langsdorf, S. Löschke, V. Möller, A. Okem,
B. Rama, eds. Cambridge University Press, Cambrige UK and New York, NY, USA. 1041–1170.
https://doi.org/10.1017/9781009325844.009.
Clark, C.M., J.G. Coughlin, J. Phelan, G. Martin, K. Austin, M. Salem, R.D. Sabo, K. Horn, R.Q.
Thomas, and R.M. Dalton. 2024. “Winners and Losers From Climate Change: An Analysis of
Climate Thresholds for Tree Growth and Survival for Roughly 150 Species Across the Contiguous
United States.” Global Change Biology 30(12):e17597. https://doi.org/10.1111/gcb.17597.
Clarke, L., J. Eom, E.H. Marten, R. Horowitz, P. Kyle, R. Link, B.K. Mignone, A. Mundra, and Y. Zhou.
2018. “Effects of Long-Term Climate Change on Global Building Energy Expenditures.” Energy
Economics 72:667–677. https://doi.org/10.1016/j.eneco.2018.01.003.
Cobb, R., and M. Metz. 2017. “Tree Diseases as a Cause and Consequence of Interacting Forest
Disturbances.” Forests 8(5):147. https://doi.org/10.3390/f8050147.
Collins, T.W., S.E. Grineski, and S.M. Nadybal. 2022. “A Comparative Approach for Environmental
Justice Analysis: Explaining Divergent Societal Distributions of Particulate Matter and Ozone
Pollution across U.S. Neighborhoods.” Annals of the American Association of Geographers
112(2):522–541. https://doi.org/10.1080/24694452.2021.1935690.
Colmer, J., I. Hardman, J. Shimshack, and J. Voorheis. 2020. “Disparities in PM2.5 Air Pollution in the
United States.” Science 369(6503):575–578. https://doi.org/10.1126/science.aaz9353.
Cook, B.I., R.L. Miller, and R. Seager. 2009. “Amplification of the North American “Dust Bowl”
Drought Through Human-Induced Land Degradation.” Proceedings of the National Academy of
Sciences 106(13):4997–5001. https://doi.org/10.1073/pnas.0810200106.
Cooper, O.R., K.-L. Chang, K. Bates, S.S. Brown, W.S. Chace, M.M. Coggon, A.M. Gorchov, et al.
2024. “Early Season 2023 Wildfires Generated Record-Breaking Surface Ozone Anomalies Across
the U.S. Upper Midwest.” Geophysical Research Letters 51(22).
https://doi.org/10.1029/2024GL111481.
Cottingham, K.L., K.C. Weathers, H.A. Ewing, M.L. Greer, and C.C. Carey. 2021. “Predicting the
Effects of Climate Change on Freshwater Cyanobacterial Blooms Requires Consideration of the
Complete Cyanobacterial Life Cycle.” Journal of Plankton Research 43(1):10–19.
https://doi.org/10.1093/plankt/fbaa059.
Couper, L.I., A.J. MacDonald, and E.A. Mordecai. 2021. “Impact of Prior and Projected Climate Change
on US Lyme Disease Incidence.” Global Change Biology 27(4):738–754.
https://doi.org/10.1111/gcb.15435.
Covert, H.H., F. Abdoel Wahid, S.E. Wenzel, and M.Y. Lichtveld. 2023. “Climate Change Impacts on
Respiratory Health: Exposure, Vulnerability, and Risk.” Physiological Reviews 103(4):2507–2522.
https://doi.org/10.1152/physrev.00043.2022.
Creed, I.F., C.M. Hewitt, N.J. Casson, A.R. Contosta, J.L. Campbell, D. Lutz, and A.T. Morzillo. 2023.
“Coupled Human-Natural System Impacts of a Winter Weather Whiplash Event.” Ecology & Society
28(2). https://doi.org/10.5751/ES-14174-280230.
Cromar, K.R., S.C. Anenberg, J.R. Balmes, A.A. Fawcett, M. Ghazipura, J.M. Gohlke, M. Hashizume, et
al. 2022. “Global Health Impacts for Economic Models of Climate Change: A Systematic Review and
Meta-Analysis.” Annals of the American Thoracic Society 19(7):1203–1212.
https://doi.org/10.1513/AnnalsATS.202110-1193OC.
Prepublication Copy
References 77
D’Amato, G., H.J. Chong‐Neto, O.P. Monge Ortega, C. Vitale, I. Ansotegui, N. Rosario, T. Haahtela, et
al. 2020. “The Effects of Climate Change on Respiratory Allergy and Asthma Induced by Pollen and
Mold Allergens. Allergy 75(9):2219–2228. https://doi.org/10.1111/all.14476.
D’Amato, G., C. Vitale, M. Lanza, A. Molino, and M. D’Amato. 2016. “Climate Change, Air Pollution,
and Allergic Respiratory Diseases: An Update.” Current Opinion in Allergy and Clinical Immunology
16(5):434–440. https://doi.org/10.1097/ACI.0000000000000301.
Dannenberg, M.P., E.K. Wise, and W.K. Smith. 2019. “Reduced Tree Growth in the Semiarid United
States Due to Asymmetric Responses to Intensifying Precipitation Extremes.” Science Advances
5(10):eaaw0667. https://doi.org/10.1126/sciadv.aaw0667.
de la Vega, E., T.B. Chalk, P.A. Wilson, R.P. Bysani, and G.L. Foster 2020. “Atmospheric CO2 During
the Mid-Piacenzian Warm Period and the M2 Glaciation.” Scientific Reports 10(1):11002.
https://doi,org/10.1038/s41598-020-67154-8.
De Roos, A.J., M.C. Kondo, L.F. Robinson, A. Rai, M. Ryan, C.N. Haas, J. Lojo, and J.A. Fagliano.
2020. “Heavy Precipitation, Drinking Water Source, and Acute Gastrointestinal Illness in
Philadelphia, 2015-2017.” PLOS ONE 15(2):e0229258.
https://doi.org/10.1371/journal.pone.0229258.
Deilami, K., M. Kamruzzaman, and Y. Liu. 2018. “Urban Heat Island Effect: A Systematic Review of
Spatio-Temporal Factors, Data, Methods, and Mitigation Measures.” International Journal of Applied
Earth Observation and Geoinformation 67:30–42. https://doi.org/10.1016/j.jag.2017.12.009
Department of Defense (DOD), Office of the Undersecretary of Defense (Acquisition and Sustainment).
2021. Department of Defense Draft Climate Adaptation Plan: Report submitted to National Climate
Task Force and Federal Chief Sustainability Officer. https://www.sustainability.gov/pdfs/dod-2021-
cap.pdf.
Depsky, N., I. Bolliger, D. Allen, J.H. Choi, M. Delgado, M. Greenstone, A. Hamidi, T. Houser, R.E.
Kopp, and S. Hsiang. 2023. “DSCIM-Coastal v1.1: An Open-Source Modeling Platform for Global
Impacts of Sea Level Rise.” Geoscientific Model Development 16(14):4331–4366.
https://doi.org/10.5194/gmd-16-4331-2023.
Deutsch, C.A., J.J. Tewksbury, M. Tigchelaar, D.S. Battisti, S.C. Merrill, R.B. Huey, and R.L. Naylor.
2018. “Increase in Crop Losses to Insect Pests in a Warming Climate.” Science 361(6405):916–919.
https://doi.org/10.1126/science.aat3466.
Dhakal, S., J.C. Minx, F.L. Toth, A. Abdel-Aziz, M.J. Figueroa Meza, K. Hubacek, I.G.C. Jonckheere, et
al. 2022. “Chapter 2: Emissions Trends and Drivers.” In Climate Change 2022: Mitigation of Climate
Change. Contribution of Working Group III to the Sixth Assessment Report of the Intergovernmental
Panel on Climate Change P.R. Shukla, J. Skea, R. Slade, A. Al Khourdajie, R. van Diemen, D.
McCollum, M. Pathak, S. Some, P. Vyas, R. Fradera, M. Belkacemi, A. Hasija, G. Lisboa, S. Luz, J.
Malley, eds. Cambridge University Press, Cambridge, UK and New York, NY, USA.
https://doi.org/10.1017/9781009157926.004.
Dharmage S.C., J.L. Perret, and A. Custovic. 2019. “Epidemiology of Asthma in Children and Adults.”
Frontiers in Pediatrics 7:246. https://doi.org/10.3389/fped.2019.00246.
Di, Q., Y. Wang, A. Zanobetti, Y. Wang, P. Koutrakis, C. Choirat, R. Dominici, and J.D. Schwartz. 2017.
“Air Pollution and Mortality in the Medicare Population.” The New England Journal of Medicine
376(26):5513-2522. https://doi.org/10.1056/NEJMoa1702747
Diaz, D.B. 2016. “Estimating Global Damages from Sea Level Rise with the Coastal Impact and
Adaptation Model (CIAM).” Climatic Change 137(1):143–156. https://doi.org/10.1007/s10584-016-
1675-4.
Dietterich, L.H., A. Zanobetti, I. Kloog, P. Huybers, A.D.B. Leakey, A.J. Bloom, E. Carlisle, et al. 2015.
“Impacts of Elevated Atmospheric CO2 on Nutrient Content of Important Food Crops.” Scientific
Data 2(150036). https://doi.org/10.1038/sdata.2015.36.
Diffenbaugh, N.S., F.V. Davenport, and M. Burke. 2021. “Historical Warming Has Increased U.S. Crop
Insurance Losses.” Environmental Research Letters 16(8):084025. https://doi.org/10.1088/1748-
9326/ac1223.
Prepublication Copy
Ditlevsen, P., and S. Ditlevsen. 2023. “Warning of a Forthcoming Collapse of the Atlantic Meridional
Overturning Circulation.” Nature Communications 14(4254). https://doi.org/10.1038/s41467-023-
39810-w.
Doney, S.C., D.S. Busch, S.R. Cooley, and K.J. Kroeker. 2020. “The Impacts of Ocean Acidification on
Marine Ecosystems and Reliant Human Communities.” Annual Review of Environment and
Resources 45(1):83–112. https://doi.org/10.1146/annurev-environ-012320-083019
Dresser, W.D., J.M. Silberstein, C.E. Reid, M.E. Vance, C. Wiedinmyer, M.P. Hannigan, and J.A. de
Gouw. 2024. “Volatile Organic Compounds Inside Homes Impacted by Smoke from the Marshall
Fire.” ACS ES&T Air 2(1):4–12. https://doi.org/10.1021/acsestair.4c00259.
Duffy, P.B., C.B. Field, N.S. Diffenbaugh, S.C. Doney, Z. Dutton, S. Goodman, L. Heinzerling, et al.
2018. “Strengthened Scientific Support for the Endangerment Finding for Atmospheric Greenhouse
Gases.” Science 363(6427). https://doi.org/10.1126/science.aat5982.
Durack, P.J. 2015. “Ocean Salinity and the Global Water Cycle.” Oceanography 28(1):20–31.
http://dx.doi.org/10.5670/oceanog.2015.03.
Durack, P.J., and S.E. Wijffels. 2010. “Fifty-Year Trends in Global Ocean Salinities and Their
Relationship to Broadscale Warming.” Journal of Climate 23:4342–4362.
https://doi.org/10.1175/2010JCLI3377.1.
Durack, P.J., S.E. Wijffels, and R.J. Matear. 2012. “Ocean Salinities Reveal Strong Global Water Cycle
Intensification during 1950 to 2000.” Science 336(6080):455–458.
https://doi.org/10.1126/science.1212222.
East, J.D., E. Monier, R.K. Saari, and F. Garcia-Menendez. 2024. “Projecting Changes in the Frequency
and Magnitude of Ozone Pollution Events under Uncertain Climate Sensitivity.” Earth’s
Future 12:e2023EF003941. https://doi.org/10.1029/2023EF003941.
Ebi, K.L., and D. Mills. 2013. “Winter Mortality in a Warming Climate: A Reassessment.” WIREs
Climate Change 4(3):203–212. https://doi.org/10.1002/wcc.211.
Ebi, K.L., and J. Nealon. 2016. “Dengue in a Changing Climate.” Environmental Research 151:115–123.
https://doi.org/10.1016/j.envres.2016.07.026.
Ebi, K.L., C. Åström, C.J. Boyer, L.J. Harrington, J.J. Hess, Y. Honda, E. Kazura, R.F. Stuart-Smith, and
F.E.L. Otto. 2020. “Using Detection and Attribution to Quantify How Climate Change Is Affecting
Health: Study Explores Detection and Attribution to Examine How Climate Change is Affecting
Health.” Health Affairs 39(12):2168–2174. https://doi.org/10.1377/hlthaff.2020.01004.
Ebi, K.L., J. Vanos, J.W. Baldwin, J.E. Bell, D.M. Hondula, N.A. Errett, K. Hayes, et al. 2021. “Extreme
Weather and Climate Change: Population Health and Health System Implications.” Annual Review of
Public Health 42(1):293–315. https://doi.org/10.1146/annurev-publhealth-012420-105026.
Eck, M.A., A.R. Murray, A.R. Ward, and C.E. Konrad. 2020. “Influence of Growing Season Temperature
and Precipitation Anomalies on Crop Yield in the Southeastern United States.” Agricultural and
Forest Meteorology 291:108053. https://doi.org/10.1016/j.agrformet.2020.108053.
Edmonds, D.A., R.L. Caldwell, E.S. Brondizio, and S.M.O. Siani. 2020 “Coastal Flooding Will
Disproportionately Impact People on River Deltas.” Nature Communications 11(1):4741.
https://doi.org/10.1038/s41467-020-18531-4
Edwards, D.A., A. Edwards, D. Li, L. Wang, K.F. Chung, D. Bhatta, A. Bilstein, et al. 2025. “Global
Warming Risks Dehydrating and Inflaming Human Airways.” Communications Earth &
Environment 6(1):1–12. https://doi.org/10.1038/s43247-025-02161-z.
Eguiluz‐Gracia, I., A.G. Mathioudakis, S. Bartel, S.J.H. Vijverberg, E. Fuertes, P. Comberiati, Y.S. Cai,
et al. 2020. “The Need for Clean Air: The Way Air Pollution and Climate Change Affect Allergic
Rhinitis and Asthma.” Allergy 75(9):2170–2184. https://doi.org/10.1111/all.14177.
Eisenman, D.P., and L.P. Galway. 2022. “The Mental Health and Well-Being Effects of Wildfire Smoke:
A Scoping Review.” BMC Public Health 22(1):2274. https://doi.org/10.1186/s12889-022-14662-z.
Ellison, A.M., M.S. Bank, B.D. Clinton, E.A. Colburn, K. Elliott, C.R. Ford, D.R. Foster, et al. 2005.
“Loss of Foundation Species: Consequences for the Structure and Dynamics of Forested
Prepublication Copy
References 79
Prepublication Copy
Erbas, B., M. Jazayeri, K.A. Lambert, C.H. Katelaris, L.A. Prendergast, R. Tham, M.J. Parrodi, et al.
2018. “Outdoor Pollen is a Trigger of Child and Adolescent Asthma Emergency Department
Presentations: A Systematic Review and Meta‐Analysis.” Allergy 73(8):1632–1641.
https://doi.org/10.1111/all.13407.
Eyring, V., S. Bony, G.A. Meehl, C.A. Senior, B. Stevens, R.J. Stouffer, and K.E. Taylor. 2016.
“Overview of the Coupled Model Intercomparison Project Phase 6 (CMIP6) Experimental Design
and Organization.” Geoscientific Model Development 9:1937–1958. https://doi.org/10.5194/gmd-9-
1937-2016.
Eyring, V., N.P. Gillett, K.M. Achuta Rao, R. Barimalala, M. Barreiro Parrillo, N. Bellouin, C. Cassou, et
al. 2021. “Human Influence on the Climate System.” In Climate Change 2021: The Physical Science
Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental
Panel on Climate Change, in Masson-Delmotte, V., P. Zhai, A. Pirani, S.L. Connors, C. Péan, S.
Berger, N. Caud, Y. Chen, L. Goldfarb, M.I. Gomis, M. Huang, K. Leitzell, E. Lonnoy, J.B.R.
Matthews, T.K. Maycock, T. Waterfield, O. Yelekçi, R. Yu, and B. Zhou eds. Cambridge University
Press, Cambridge, United Kingdom and New York, NY, USA. 423–552.
https://doi.org/10.107/9781009157896.005
Fahey, T.J., T.G. Siccama, C.T. Driscoll, G.E. Likens, J. Campbell, C.E. Johnson, J.J. Battles, et al. 2005.
“The Biogeochemistry of Carbon at Hubbard Brook.” Biogeochemistry 75(1):109–176.
https://doi.org/10.1007/s10533-004-6321-y.
Fall, S., K.M. Coulibaly, J.E. Quansah, G. El Afandi, and R. Ankumah. 2021. “Observed Daily
Temperature Variability and Extremes over Southeastern USA (1978–2017).” Climate 9:110.
https://doi.org/10.3390/cli9070110.
Fann, N.L., C.G. Nolte, M.C. Sarofim, J. Martinich, and N.J. Nassikas. 2021. “Associations Between
Simulated Future Changes in Climate, Air Quality, and Human Health.” JAMA Network
4(1):e2032064. https://doi.org/10.1001/jamanetworkopen.2020.32064.
Fedewa, E.J., T.M. Jackson, J.I. Richar, J.L. Gardner, and M.A. Litzow. 2020. “Recent Shifts in Northern
Bering Sea Snow Crab (Chionoecetes opilio) Size Structure and the Potential Role of Climate-
Mediated Range Contraction.” Deep Sea Research Part II: Topical Studies in Oceanography 181–
182:104878. https://doi.org/10.1016/j.dsr2.2020.104878.
Feng, Y., H. Zhang, S. Tao, Z. Ao, C. Song, J. Chave, T. Le Toan, et al. 2022. “Decadal Lake Volume
Changes (2003–2020) and Driving Forces at a Global Scale.” Remote Sensing 14(4):1032.
https://doi.org/10.3390/rs14041032.
Fetterer, F., K. Knowles, W.N. Meier, M. Savoie, A.K. Windnagel, and T. Stafford. 2025. Sea Ice Index.
(G02135, Version 4). [Data Set]. Boulder, Colorado USA. National Snow and Ice Data
Center. https://doi.org/10.7265/a98x-0f50.
Fiore, A.M., D.J. Jacob, B.D. Field, D.G. Streets, S.D. Fernandes, and C. Jang. 2002. “Linking Ozone
Pollution and Climate Change: The Case for Controlling Methane.” Geophysical Research Letters
29(19). https://doi.org/10.1029/2002gl015601.
Fiore, A.M., J.J. West, L.W. Horowitz, V. Naik, and M.D. Schwarzkopf. 2008. “Characterizing the
Tropospheric Ozone Response to Methane Emission Controls and the Benefits to Climate and Air
Quality.” Journal of Geophysical Research: Atmospheres 113(D8).
https://doi.org/10.1029/2007JD009162.
Fischer, E.M., M. Bador, R. Huser, E.J. Kendon, A. Robinson, and S. Sippel. 2025. “Record-Breaking
Extremes in a Warming Climate.” Nature Reviews: Earth & Environment 6:456–470.
https://doi.org/10.1038/s43017-025-00681-y.
Fisher, M.C., S.K. Moore, S.L. Jardine, J.R. Watson, and J.F. Samhouri. 2021. “Climate Shock Effects
and Mediation in Fisheries.” Proceedings of the National Academy of Sciences of the United States of
America 118(2):e2014379117. https://doi.org/10.1073/pnas.2014379117.
Flouris, A.D., P.C. Dinas, L.G. Ioannou, L. Nybo, G. Havenith, G.P. Kenny, and T. Kjellstrom. 2018.
“Workers’ Health and Productivity under Occupational Heat Strain: A Systematic Review and Meta-
Prepublication Copy
References 81
Prepublication Copy
Prepublication Copy
References 83
Gwon, Y., Y. Ji, J.D. Berman, A.M. Abadi, R.D. Leeper, J. Rennie, and J.E. Bell. 2025. “Impacts of
Drought on Respiratory Mortality in the Upper Midwest United States: A Population Subgroup
Assessment.” Environmental Research: Health 3(2):025002. https://doi.org/10.1088/2752-
5309/adafd6.
Hahn, M.B., A.J. Monaghan, M.H. Hayden, R.J. Eisen, M.J. Delorey, N.P. Lindsey, R.S. Nasci, and M.
Fischer. 2015. “Meteorological Conditions Associated with Increased Incidence of West Nile Virus
Disease in the United States, 2004–2012.” The American Society of Tropical Medicine and Hygiene
92(5):1013–1022. https://doi.org/10.4269/ajtmh.14-0737.
Hakuba, M.Z., S. Fourest, T. Boyer, B. Meyssignac, J.A. Carton, G. Forget, L. Cheng, et al.
2024. “Trends and Variability in Earth’s Energy Imbalance and Ocean Heat Uptake Since
2005.” Surveys in Geophysics 45:1721–1756. https://doi.org/10.1007/s10712-024-09849-5.
Haley, B.M., Y. Sun, J.S. Jagai, J.H. Leibler, R. Fulweiler, J. Ashmore, G.A. Wellenius, and W. Heiger-
Bernays. 2024. “Association between Combined Sewer Overflow Events and Gastrointestinal Illness
in Massachusetts Municipalities with and without River-Sourced Drinking Water, 2014–2019.”
Environmental Health Perspectives 132(5):057008. https://doi.org/10.1289/EHP14213.
Hall, D.K., J.S. Kimball, R. Larson, N.E. DiGirolamo, K.A. Casey, and G. Hulley. 2023. “Intensified
Warming and Aridity Accelerate Terminal Lake Desiccation in the Great Basin of the Western United
States.” Earth and Space Science 10:e2022EA002630. https://doi.org/10.1029/2022EA002630
Ham, Y.-G., J.-H. Kim, S.-K. Min, D. Kim, T. Li, A. Timmermann, and M.F. Stuecker. 2023.
“Anthropogenic Fingerprints in Daily Precipitation Revealed by Deep Learning.” Nature 622:301–
307. https://doi.org/10/1038/s14586-023-06474-x
Hamlington, B.D., A. Bellas-Manley, J.K. Willis, S. Fournier, N. Vinogradova, R.S. Nerem, P.R.
Thompson, and R. Kopp. 2024. “The Rate of Global Sea Level Rise Doubled During the Past Three
Decades.” Communications Earth & Environment 5(1):601. https://doi.org/10.1038/s43247-024-
01761-5.
Hamlington, B.D., D.P. Chambers, T. Frederikse, S. Dangendorf, S. Fournier, B. Buzzanga, and R.S.
Nerem. 2022. “Observation-Based Trajectory of Future Sea Level for the Coastal United States
Tracks Near High-End Model Projections.” Communications Earth Environment 3(230).
https://doi.org/10.1038/s43247-022-00537-z.
Hammond, J.C., C. Simeone, J.S. Hecht, G.A. Hodgkins, M. Lombard, G. McCabe, D. Wolock,, et al.
2022. “Going Beyond Low Flows: Streamflow Drought Deficit and Duration Illuminate Distinct
Spatiotemporal Drought Patterns and Trends in the U.S. During the Last Century.” Water Resources
Research 58(9):e2022WR031930. https://doi.org/10.1029/2022WR031930
Han, Z., and H.O. Sharif. 2021. “Analysis of Flood Fatalities in the United States, 1959–2019.” Water
13(13):1871. https://doi.org/10.3390/w13131871.
Hansson, E., K. Broberg, J. Wijkström, J. Glaser, M. Gonzalez-Quiroz, U. Ekström, M. Abrahamson, and
K. Jakobsson. 2023. “An Explorative Study of Inflammation-Related Proteins Associated with
Kidney Injury in Male Heat-Stressed Workers.” Journal of Thermal Biology 112:103433.
https://doi.org/10.1016/j.jtherbio.2022.103433.
Hao, Y., L. Balluz, H. Strosnider, X.J. Wen, C. Li, and J.R. Qualters. 2015. “Ozone, Fine Particulate
Matter, and Chronic Lower Respiratory Disease Mortality in the United States.” American Journal of
Respiratory and Critical Care Medicine 192(3):337–341. https://doi.org/10.1164/rccm.201410-
1852OC.
Harries, J.E., H.E. Brindley, P.J. Sagoo, and R. Bantges. 2001. “Increases in Greenhouse Forcing Inferred
from the Outgoing Longwave Radiation Spectra of the Earth in 1970 and 1997. Nature
410(6826):355–357. https://doi.org/10.1038/35066553.
Hartin, C., E.E. McDuffie, K. Noiva, M. Sarofim, B. Parthum, J. Martinich, S. Barr, J. Neumann, J.
Willwerth, and A. Fawcett. 2023. “Advancing the Estimation of Future Climate Impacts Within the
United States.” Earth System Dynamics 14(5):1015–1037. https://doi.org/10.5194/esd-14-1015-2023.
Prepublication Copy
Hatfield, J.L., and J.H. Prueger. 2015. “Temperature Extremes: Effect on Plant Growth and
Development.” Weather and Climate Extremes 10(A):4–10.
https://doi.org/10.1016/j.wace.2015.08.001
Hatfield, J.L., K.J. Boote, B.A. Kimball, L.H. Ziska, R.C. Izaurralde, D. Ort, A.M. Thomson, and D.
Wolfe. 2011. “Climate Impacts on Agriculture: Implications for Crop Production.” Agronomy
Journal 103(2):351–370. https://doi.org/10.2134/agronj2010.0303.
Hausfather, Z. 2025. “An Assessment of Current Policy Scenarios over the 21st Century and the Reduced
Plausibility of High-Emissions Pathways.” Dialogues on Climate Change 2(1):26–32.
https://doi.org/10.1177/29768659241304854
Hausfather, Z., H.F. Drake, T. Abbott, and G.A. Schmidt, 2020. “Evaluating the Performance of Past
Climate Model Projections.” Geophysical Research Letters 47(1):e2019GL085378.
https://doi.org/10.1029/2019GL085378.
Hawkins, E., and R. Sutton. 2009. “The Potential to Narrow Uncertainty in Regional Climate
Predictions.” Bulletin of the American Meteorological Society 90:1095–1108.
https://doi.org/10.1175/2009BAMS2607.1.
Hayden, M.H., P.J. Schramm, C.B. Beard, J.E. Bell, A.S. Bernstein, A. Bieniek-Tobasco, N. Cooley, et
al. 2023. “Chapter 15. Human Health.” In Fifth National Climate Assessment. Crimmins, A.R., C.W.
Avery, D.R. Easterling, K.E. Kunkel, B.C. Stewart, and T.K. Maycock, eds. U.S. Global Change
Research Program, Washington, DC, USA. https://doi.org/10.7930/NCA5.2023.CH15.
He, C., S. Breitner-Busch, V. Huber, K. Chen, S. Zhang, A. Gasparrini, M. Bell, et al. 2024. “Rainfall
Events and Daily Mortality Across 645 Global Locations: Two Stage Time Series Analysis.” BMJ
387:e080944. https://doi.org/10.1136/bmj-2024-080944
He, Q., J. Cao, P.E. Saide, T. Ye, W. Wang, M. Zhang, and J. Huang. 2025. “Evaluating Spatiotemporal
Variations and Exposure Risk of Ground-Level Ozone Concentrations Across China from 2000 to
2020 Using Satellite-Derived High-Resolution Data.” Atmospheric Chemistry and Physics
25(13)6663–6677. https://doi.org/10.5194/acp-25-6663-2025.
Head, J.R., G. Sondermeyer-Cooksey, A.K. Heaney, A.T. Yu, I. Jones, A. Bhattachan, S.K. Campo, et al.
2022. “Effects of Precipitation, Heat, and Drought on Incidence and Expansion of
Coccidioidomycosis in Western USA: A Longitudinal Surveillance Study.” The Lancet Planetary
Health 6(10):793–803. https://doi.org/10.1016/S2542-5196(22)00202-9.
Healy, J.P., M. Danesh Yazdi, Y. Wei, X. Qiu, A. Shtein, F. Dominici, L. Shi, and J.D. Schwartz. 2023.
“Seasonal Temperature Variability and Mortality in the Medicare Population.” Environmental Health
Perspectives 131(7):077002. https://doi.org/10.1289/EHP11588.
Heft-Neal, S., A. Driscoll, W. Yang, G. Shaw, and M. Burke. 2022. “Associations Between Wildfire
Smoke Exposure During Pregnancy and Risk of Preterm Birth in California.” Environmental
Research 203:111872. https://doi.org/10.1016/j.envres.2021.111872
Hegerl, G.C., H. von Storch, K. Hasselmann, B.D. Santer, U. Cubasch, and P.D. Jones. 1996. “Detecting
Greenhouse Gas Induced Climate Change with an Optimal Fingerprint Method.” Journal of
Climate 9:2281–2306. https://doi.org/10.1175/1520-0442(1996)009<2281:DGGICC>2.0.CO;2
Heilmann, A., Z. Rueda, D. Alexander, K.B. Laupland, and Y. Keynan. 2024. “Impact of Climate Change
on Amoeba and the Bacteria They Host.” Journal of the Association of Medical Microbiology and
Infectious Disease Canada 9(1)1–5. https://doi.org/10.3138/jammi-2023-09-08.
Held, I.M., and B.J. Soden. 2006. “Robust Responses of the Hydrological Cycle to Global Warming.”
Journal of Climate 19:5686–5699. https://doi.org/10.1175/JCLI3990.1
Henny, L., and K. Kim. 2025. “The Changing Nature of Atmospheric Rivers.” Journal of
Climate 38:1435–1456. https://doi.org/10.1175/JCLI-D-24-0234.1.
Hepler, S.A., K.A. Kaufeld, K. Benedict, M. Toda, B.R. Jackson, X. Liu, and D. Kline. 2022. “Integrating
Public Health Surveillance and Environmental Data to Model Presence of Histoplasma in the United
States.” Epidemiology 33(5)654–659. https://doi.org/10.1097/EDE.0000000000001499.
Prepublication Copy
References 85
Herrera, D.A., B.I. Cook, J. Fasullo, K.J. Anchukaitis, M. Alessi, C.J. Martinez, C.P. Evans, et al. 2023.
“Observed Changes in Hydroclimate Attributed to Human Forcing.” PLOS Climate 2(11):e0000303.
https://doi.org/10.1371/journal.pclm.0000303.
Hicke, J.A., S. Lucatello, L.D. Mortsch, J. Dawson, M. Domínguez Aguilar, C.A.F. Enquist, E.A.
Gilmore, et al. 2022. “North America.” In Climate Change 2022: Impacts, Adaptation and
Vulnerability. Contribution of Working Group II to the Sixth Assessment Report of the
Intergovernmental Panel on Climate Change H.-O. Pörtner, D.C. Roberts, M. Tignor, E.S.
Poloczanska, K. Mintenbeck, A. Alegría, M. Craig, S. Langsdorf, S. Löschke, V. Möller, A. Okem,
B. Rama, eds. Cambridge University Press, Cambridge, UK and New York, NY, USA. 1929–2042.
https://doi.org/10.1017/9781009325844.016.
Hickman, C., E. Marks, P. Pihkala, S. Clayton, R.E. Lewandowski, E.E. Mayall, B. Wray, C. Mellor, and
L. van Susteren. 2021. “Climate Anxiety in Children and Young People and Their Beliefs about
Government Responses to Climate Change: A Global Survey.” The Lancet Planetary Health
5(12):e863–e873. https://doi.org/10.1016/S2542-5196(21)00278-3.
Higuera, P.E., M.C. Cook, J.K. Balch, E.N. Stavros, A.L. Mahood, and L.A. St. Denis. 2023. “Shifting
Social-Ecological Fire Regimes Explain Increasing Structure Loss from Western Wildfires.” PNAS
Nexus 2(3):pgad005. https://doi.org/10.1093/pnasnexus/pgad005.
Hoare, E., F. Jacka, and M. Berk. 2019. “The Impact of Urbanization on Mood Disorders: An Update of
Recent Evidence.” Current Opinion in Psychiatry 32(3):198–203.
https://doi.org/10.1097/YCO.0000000000000487.
Hobbs, W., P. Spence, A. Meyer, S. Schroeter, A.D. Fraser, P. Reid, T.R. Tian, et al. 2024.
“Observational Evidence for a Regime Shift in Summer Antarctic Sea Ice.” Journal of
Climate 37(7):2263–2275. https://doi.org/10.1175/JCLI-D-23-0479.1
Hogan, J.A., G.M. Domke, K. Zhu, D.J. Johnson, and J.W. Lichstein. 2024. “Climate Change Determines
the Sign of Productivity Trends in US Forests.” Proceedings of the National Academy of Sciences
121(4):e2311132121. https://doi.org/10.1073/pnas.2311132121.
Holm, S.M., and J.R. Balmes. 2022. “Systematic Review of Ozone Effects on Human Lung Function,
2013 Through 2020.” CHEST 161(1):190–201. https://doi.org/10.1016/j.chest.2021.07.2170.
Howard, J.T., N. Androne, K.C. Alcover, and A.R. Santos-Lozada. 2024. “Trends of Heat-Related Deaths
in the US, 1999-2023. JAMA 332(14):1203–1204. https://doi.org/10.1001/jama.2024.16386.
Howard, M.H., C.M. Sayes, J.P. Giesy, and Y. Li. 2024a. “Valley Fever Under a Changing Climate in the
United States.” Environment International 193:109066. https://doi.org/10.1016/j.envint.2024.109066.
Howe, P.D., J.R. Marlon, M. Mildenberger, and B.S. Shield. 2019. “How Will Climate Change Shape
Climate Opinion?” Environmental Research Letters 14:113001. https://doi.org/10.1088/1748-
9326/ab466a.
Huang, H., M.R. Fischella, Y. Liu, Z. Ban, J.V. Fayne, D. Li, K.C. Cavanaugh, and D.P. Lettenmaier.
2022. “Changes in Mechanisms and Characteristics of Western U.S. Floods Over the Last Sixty
Years.” Geophysical Research Letters 49(3):e2021GL097022.
https://doi.org/10.1029/2021GL097022.
Huntington, H.P., C. Strawhacker, J. Falke, E.M. Ward, L. Behnken, T.N. Curry, A.C. Herrmann, et al.
2023. “Chapter 29. Alaska.” In Fifth National Climate Assessment. Crimmins, A.R., C.W. Avery,
D.R. Easterling, K.E. Kunkel, B.C. Stewart, and T.K. Maycock, eds. U.S. Global Change Research
Program, Washington, DC, USA. https://doi.org/10.7930/NCA5.2023.CH29.
Hutchins, J., M. Skidmore, and D. Nolan. 2025. “Vulnerability of US Dairy Farms to Extreme Heat.”
Food Policy 131:102821. https://doi.org/10.1016/j.foodpol.2025.102821.
Ibebuchi, C.C., C.C. Lee, and S.C. Sheridan. 2025. “Recent Trends in Extreme Temperature Events
Across the Contiguous United States. International Journal of Climatology
45:e8693. https://doi.org/10.1002/joc.8693.
Idrose, N.S., C.J. Lodge, B. Erbas, J.A. Douglass, D.S. Bui, and S.C. Dharmage. 2022. “A Review of the
Respiratory Health Burden Attributable to Short-Term Exposure to Pollen.” International Journal of
Environmental Research and Public Health 19(12):7541. https://doi.org/10.3390/ijerph19127541.
Prepublication Copy
Intergovernmental Panel on Climate Change (IPCC). 2022. Climate Change 2022: Mitigation of Climate
Change. Contribution of Working Group III to the Sixth Assessment Report of the Intergovernmental
Panel on Climate Change [P.R. Shukla, J. Skea, R. Slade, A. Al Khourdajie, R. van Diemen, D.
McCollum, M. Pathak, S. Some, P. Vyas, R. Fradera, M. Belkacemi, A. Hasija, G. Lisboa, S. Luz, J.
Malley, (eds.)]. Cambridge University Press, Cambridge, UK and New York, NY, USA.
https://doi.org/10.1017/9781009157926.
IPCC. 2022a. “Climate Change 2022: Impacts, Adaptation and Vulnerability.” Contribution of Working
Group II to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. H.-O.
Pörtner, D.C. Roberts, M. Tignor, E. S. Poloczanska, K. Mintenbeck, A. Alegría, M. Craig, S.
Langsdorf, S. Löschke, V. Möller, A. Okem, and B. Rama, eds. Cambridge, UK and New York, NY,
USA: Cambridge University Press. http://doi.org/10.1017/9781009325844.
IPCC. 2021. “Climate Change 2021: The Physical Science Basis.” Contribution of Working Group I to
the Sixth Assessment Report of the Intergovernmental Panel on Climate Change Masson-Delmotte,
V., P. Zhai, A. Pirani, S.L. Connors, C. Péan, S. Berger, N. Caud, Y. Chen, L. Goldfarb, M.I. Gomis,
M. Huang, K. Leitzell, E. Lonnoy, J.B.R. Matthews, T.K. Maycock, T. Waterfield, O. Yelekçi, R. Yu,
and B. Zhou, eds. Cambridge University Press, Cambridge, United Kingdom and New York, NY,
USA. https://doi.org/10.1017/9781009157896.
IPCC. 2019. “IPCC Special Report on the Ocean and Cryosphere in a Changing Climate.” [H.-O. Pörtner,
D.C. Roberts, V. Masson-Delmotte, P. Zhai, M. Tignor, E. Poloczanska, K. Mintenbeck, A. Alegría,
M. Nicolai, A. Okem, J. Petzold, B. Rama, N.M. Weyer (eds.)]. Cambridge University Press,
Cambridge, UK and New York, NY, USA. https://doi.org/10.1017/9781009157964.
IPCC. 2007. Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working
Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. M.L.
Parry, O.F. Canziani, J.P. Palutikof, P.J. van der Linden, and C.E. Hanson, eds. Cambridge University
Press, Cambridge, UK. https://www.ipcc.ch/report/ar4/wg2.
Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services (IPBES). 2019.
“Global Assessment Report on Biodiversity and Ecosystem Services of the Intergovernmental
Science-Policy Platform on Biodivsersity and Ecosystem Service.” IPBES. Brondizio, E., D. Sandra,
J. Settele, and H.T. Ngo, eds.
Ioannou, L.G., J. Foster, N.B. Morris, J.F. Piil, G. Havenith, I.B. Mekjavic, G.P. Kenny, L. Nybo, and
A.D. Flouris. 2022. “Occupational Heat Strain in Outdoor Workers: A Comprehensive Review and
Meta-Analysis.” Temperature 9(1):67–102. https://doi.org/10.1080/23328940.2022.2030634.
Irrgang, A.M., M. Bendixen, L.M. Farquharson, A.V. Baranskaya, L.H. Erikson, A.E. Gibbs, S.A.
Ogorodov, et al. 2022. “Drivers, Dynamics and Impacts of Changing Arctic Coasts.” Nature Reviews
Earth & Environment 3(1):39–54. https://doi.org/10.1038/s43017-021-00232-1.
Jackson, L.C., E. Alastrué de Asenjo, K. Bellomo, G. Danabasoglu, H. Haak, A. Hu, J. Jungclaus, et al.
2023. “Understanding AMOC Stability: The North Atlantic Hosing Model Intercomparison Project”
Geoscientific Model Development 16:1975–1995. https://doi.org/10.5194/gmd-16-1975-2023.
Jane, S.F., T.M. Detmer, S.L. Larrick, K.C. Rose, E.A. Randall, K.J. Jirka, and P.B. McIntyre. 2024.
“Concurrent Warming and Browning Eliminate Cold-Water Fish Habitat in Many Temperate Lakes.”
Proceedings of the National Academy of Sciences 121(2):e2306906120.
https://doi.org/10.1073/pnas.2306906120.
Jane, S.F., G.J.A. Hansen, B.M. Kraemer, P.R. Leavitt, J.L. Mincer, R.L. North, R.M. Pilla, et al. 2021.
“Widespread Deoxygenation of Temperate Lakes.” Nature 594(7861):66–70.
https://doi.org/10.1038/s41586-021-03550-y.
Jasinski, M.F., J.S. Borak, S.V. Kumar, D.M. Mocko, C.D. Peters-Lidard, M. Rodell, H. Rui, et al. 2019.
“NCA-LDAs: Overview and Analysis of Hydrologic Trends for National Climate Assessment.”
Journal of Hydrometeorology 20(8):1595–11617. https://doi.org/10.1175/jhm-d-17-0234.1
Jbaily, A., X. Zhou, J. Liu, T.-H. Lee, L. Kamareddine, S. Verguet, and F. Dominici. 2022. “Air Pollution
Exposure Disparities Across US Population and Income Groups.” Nature 601(7892):228–233.
https://doi.org/10.1038/s41586-021-04190-y
Prepublication Copy
References 87
Jeevanjee, N., D.J. Paynter, J.P. Dunne, L.T. Sentman, J.P. Krasting. 2025. “A Holistic View of Climate
Sensitivity.” Annual Review of Earth and Planetary Sciences 53:367–396.
https://doi.org/10.1146/annurev-earth-040523-014302.
Jerrett, M., R.T. Burnett, C.A. Pope, K. Ito, G. Thurston, D. Krewski, Y. Shi, E. Calle, and M. Thun.
2009. “Long-Term Ozone Exposure and Mortality.” New England Journal of Medicine
360(11):1085–1095. https://doi.org/10.1056/NEJMoa0803894
Jiao, A., Y. Sun, C. Avila, V. Chiu, J. Slezak, D.A. Sacks, J.T. Abatzoglou, et al. 2023. “Analysis of Heat
Exposure During Pregnancy and Severe Maternal Morbidity.” JAMA Network Open 6(9):e2332780.
https://doi.org/10.1001/jamanetworkopen.2023.32780.
Jimenez, M.P., N.V. DeVille, E.G. Elliott, J.E. Schiff, G.E. Wilt, J.E. Hart, and P. James. 2021.
“Associations between Nature Exposure and Health: A Review of the Evidence.” International
Journal of Environmental Research and Public Health 18(9):4790.
https://doi.org/10.3390/ijerph18094790.
Jin, D., S. Moore, D. Holland, L. Anderson, W.A. Lim, D. Kim, S. Jardine, S. Martino, F. Gainella, and
K. Davidson. 2020. “2 Evaluating the Economic Impacts of Harmful Algal Blooms: Issues, Methods,
and Examples.” PICES Scientific Report (59):5-41.
https://repository.oceanbestpractices.org/bitstream/handle/11329/1850/Rpt59.pdf?sequence=1&isAllo
wed=y#page=10.
Jin, Z., G.A. Ferrada, D. Zhang, N. Scovronick, J.S. Fu, K. Chen, and Y. Liu. 2025. “Fire Smoke
Elevated the Carbonaceous PM2.5 Concentration and Mortality Burden in the Contiguous U.S. and
Southern Canada.” Environmental Science & Technology 59(24):12196–12210.
https://doi.org/10.1021/acs.est.5c01641.
Johnson, R.J., C. Wesseling, and L.S. Newman. 2019. “Chronic Kidney Disease of Unknown Cause in
Agricultural Communities.” New England Journal of Medicine 380(19):1843–1852.
https://doi.org/10.1056/NEJMra1813869
Jones, B.M., A.M. Irrgang, L.M. Farquharson, H. Lantuit, D. Whalen, S. Ogorodov, M. Grigoriev, et al.
2020. NOAA Arctic Report Card 2020: Coastal Permafrost Erosion. https://doi.org/10.25923/E47W-
DW52.
Jones, C.D., T.L. Frölicher, C. Koven, A.H. MacDougall, H.D. Matthews, K. Zickfeld, J. Rogelj, et al.
2019. “The Zero Emissions Commitment Model Intercomparison Project (ZECMIP) Contribution to
C4MIP: Quantifying Committed Climate Changes Following Zero Carbon Emissions.” Geoscientific
Model Development 12:4375–4385. https://doi.org/10.5194/gmd-12-4375-2019.
Jones, J.A., and C.T. Driscoll. 2022. “Long-Term Ecological Research on Ecosystem Responses to
Climate Change.” BioScience 72(9):814–826. https://doi.org/10.1093/biosci/biac021.
Jones, M.E., T.D. Paine, M.E. Fenn, and M.A. Poth. 2004. “Influence of Ozone and Nitrogen Deposition
on Bark Beetle Activity under Drought Conditions.” Forest Ecology and Management 200(1–3):67–
76. https://doi.org/10.1016/j.foreco.2004.06.003.
Jones, M.W., G.P. Peters, T. Gasser, R.M. Andrew, C. Schwingshackl, J. Gütschow, R.A. Houghton, P.
Friedlingstein, J. Pongratz, and C. Le Quéré. 2024a. “National Contributions to Climate Change due
to Historical Emissions of Carbon Dioxide, Methane and Nitrous Oxide.” Scientific Data 10(155).
https://doi.org/10.5281/zenodo.14054503
Jones, M.W., S. Veraverbeke, N. Andela, S.H. Doerr, C. Kolden, G. Mataveli, M.L. Pettinari, et al.
2024b. “Global Rise in Forest Fire Emissions Linked to Climate Change in the Extratropics.” Science
386(6719). https://doi.org/10.1126/science.adl5889.
Jones‐Ngo, C.G., R.J. Schmidt, E. Monier, S. Ludwick, M.Z. Al‐Hamdan, J. Vargo, and K.C. Conlon.
2025. “Joint Effects of Wildfire Smoke and Extreme Heat on Hospitalizations in California, 2011–
2020.” GeoHealth 9(6):e2024GH001237. 10.1029/2024GH001237
Jonkman, S.N., A. Curran, and L.M. Bouwer. 2024. “Floods Have Become Less Deadly: An Analysis of
Global Flood Fatalities 1975–2022.” Natural Hazards 120:632706342.
https://doi.org/10.1007/s11069-024-06444-0
Prepublication Copy
Kalashnikov, D.A., J.L. Schnell, J.T. Abatzoglou, D.L. Swain, and D. Singh. 2022. “Increasing Co-
Occurrence of Fine Particulate Matter and Ground-Level Ozone Extremes in the Western United
States.” Science Advances 8. https://doi.org/10.1126/sciadv.abi9386
Kalkuhl, M., and L. Wenz. 2020. “The Impact of Climate Conditions on Economic Production. Evidence
from a Global Panel of Regions.” Journal of Environmental Economics and Management
103:102360. https://doi.org/10.1016/j.jeem.2020.102360
Kay, J.E., C. Deser, A. Phillips, A. Mai, C. Hannay, G. Strand, J.M. Arblaster, et al. 2015. “The
Community Earth System Model (CESM) Large Ensemble Project: A Community Resource for
Studying Climate Change in the Presence of Internal Climate Variability.” Bulletin of the American
Meteorological Society 96(8):1333–1349. https://doi.org/10.1175/BAMS-D-13-00255.1.
Kay, J., M.M. Holland, and A. Jahn. 2011. “Inter-Annual to Multi-Decadal Arctic Sea Ice Extent Trends
in a Warming World.” Geophysical Research Letters 38(15):L15708.
https://doi.org/10.1029/2011GL048008.
Kazemiparkouhi, F., K.-D. Eum, B. Wang, J. Manjourides, and H.H. Suh. 2020. “Long-Term Ozone
Exposures and Cause-Specific Mortality in a US Medicare Cohort.” Journal of Exposure Science &
Environmental Epidemiology 30(4):650–658. https://doi.org/10.1038/s41370-019-0135-4
Kazi, D.S., E. Katznelson, C.L. Liu, N.M. Al-Roub, R.S. Chaudhary, D.E. Young, M. McNichol, et al.
2024. “Climate Change and Cardiovascular Health: A Systematic Review.” JAMA Cardiology
9(8):748–757. https://doi.org/10.1001/jamacardio.2024.1321.
Keeling, C.D., S.C. Piper, R.B. Bacastow, M. Wahlen, T.P. Whorf, M. Heimann, and H.A. Meijer. 2001.
“Exchanges of Atmospheric CO2 and 13CO2 with the Terrestrial Biosphere and Oceans from 1978 to
2000: Observations and Carbon Cycle Implications.” SIO Reference Series 01-06:83-113. Scripps
Institution of Oceanography, San Diego. http://escholarship.org/uc/item/09v319r9
Kemter, M., N. Marwan, G. Villarini, and B. Merz. 2023. “Controls on Flood Trends Across the United
States.” Water Resources Research 59(2):e2021WR031673. https://doi.org/10.1029/2021WR031673
Key, N., S. Sneeringer, and D. Marquardt. 2014. “Climate Change, Heat Stress, and U.S. Dairy
Production.” ERR-175, U.S. Department of Agriculture, Economic Research Service.
Khalili, R., Y. Liu, Y. Xu, K. O’Sharkey, N. Pavlovic, C. McClure, F. Lurmann, et al. 2025. “Adverse
Birth Outcomes Associated with Heat Stress and Wildfire Smoke Exposure During Preconception
and Pregnancy.” Environmental Science & Technology 59(25):12458–12471.
https://doi.org/10.1021/acs.est.4c10194
Khatana, S.A.M., J.J. Szeto, L.A. Eberly, A.S. Nathan, J. Puvvula, and A. Chen. 2024. “Projections of
Extreme Temperature–Related Deaths in the US.” JAMA Network Open 7(9):e2434942.
https://doi.org/10.1001/jamanetworkopen.2024.34942.
Khatana, S.A.M., R.M. Werner, and P.W. Groeneveld. 2022. “Association of Extreme Heat and
Cardiovascular Mortality in the United States: A County-Level Longitudinal Analysis From 2008 to
2017.” Circulation 146(3):249–261. https://doi.org/10.1161/CIRCULATIONAHA.122.060746.
Kim, Y., C.F.S. Ng, Y. Chung, H. Kim, Y. Honda, Y.L. Guo, Y.-H. Lim, B.-Y. Chen, L.A. Page, and M.
Hashizume. 2018. “Air Pollution and Suicide in 10 Cities in Northeast Asia: A Time-Stratified Case-
Crossover Analysis.” Environmental Health Perspectives 126(3):037002.
https://doi.org/10.1289/EHP2223.
Kimball, B.A. 2010. “Lessons from FACE: CO2 Effects and Interactions with Water, Nitrogen and
Temperature.” In D. Hillel and C. Rosenzweig, Series on Climate Change Impacts, Adaptation, and
Mitigation 1:87–107. Imperial College Press. https://doi.org/10.1142/9781848166561_0006.
Kirchmeier-Young, M.C., and X. Zhang. 2020. “Human Influence has Intensified Extreme Precipitation
in North America.” Proceedings of the National Academy of Sciences 117(24):13308–13313.
https://doi.org/10.1073/pnas.1921628117.
Klepeis, N.E., W.C. Nelson, W.R. Ott, J.P. Robinson, A.M. Tsang, P. Switzer, J.V. Behar, S.C. Hern, and
W.H. Engelmann. 2001. “The National Human Activity Pattern Survey (NHAPS): A Resource for
Assessing Exposure to Environmental Pollutants.” Journal of Exposure Science & Environmental
Epidemiology 11(3):231–252. https://doi.org/10.1038/sj.jea.7500165.
Prepublication Copy
References 89
Klompmaker, J.O., F. Laden, P. James, M.B. Sabath, X. Wu, J. Schwartz, F. Dominici, A. Zanobetti, and
J.E. Hart. 2023. “Effects of Long-Term Average Temperature on Cardiovascular Disease
Hospitalizations in an American Elderly Population.” Environmental Research 216:114684.
https://doi.org/10.1016/j.envres.2022.114684.
Knutson, T.R., M.V. Chung, G. Vecchi, J. Sun, T.-L. Hsieh, and A.J.P. Smith. 2021. “Climate Change is
Probably Increasing the Intensity of Tropical Cyclones.” In: Critical Issues in Climate Change
Science: ScienceBrief, C. Le Quéré, P. Liss, and P. Forster, eds.
https://doi.org/10.5281/zenodo.4570334.
Koerner, S.E., M.L. Avolio, J.M. Blair, A.K. Knapp, and M.D. Smith. 2023. “Multiple Global Change
Drivers Show Independent, Not Interactive Effects: A Long-
Term Case Study in Tallgrass Prairie.” Oecologia 201(1):143–154. https://doi.org/10.1007/s00442-
022-05295-5
Kossin, J.P. 2018. “A Global Slowdown of Tropical-Cyclone Translation Speed.” Nature 558(7708):104–
107. https://doi.org/10.1038/s41586-018-0158-3.
Kossin, J.P., K.R. Knapp, T.L. Olander, and C.S. Velden. 2020. “Global Increase in Major Tropical
Cyclone Exceedance Probability over the Past Four Decades.” Proceedings of the National Academy
of Sciences 117(22):11975–11980. https://doi.org/10.1073/pnas.1920849117.
Kotz, M., A. Levermann, and L. Wenz. 2022. “The Effect of Rainfall Changes on Economic Production.”
Nature 601(7892):223–227. https://doi.org/10.1038/s41586-021-04283-8
Kotz, M., L. Wenz, A. Stechemesser, M. Kalkuhl, and A. Levermann. 2021. “Day-to-Day Temperature
Variability Reduces Economic Growth.” Nature Climate Change 11(4):319–325.
https://doi.org/10.1038/s41558-020-00985-5
Krebs, B., and M. Neidell. 2024. “Wildfires Exacerbate Inequalities in Indoor Pollution Exposure.”
Environmental Research Letters 19(2):024043. https://doi.org/10.1088/1748-9326/ad22b8.
Kuehn, L., and S. McCormick. 2017. “Heat Exposure and Maternal Health in the Face of Climate
Change.” International Journal of Environmental Research and Public Health 14(8):853.
https://doi.org/10.3390/ijerph14080853
Kugeler K.J., A. Earley, P.S. Mead, and A.F. Hinckley. 2024. “Surveillance for Lyme Disease After
Implementation of a Revised Case Definition — United States, 2022.” Morbidity and Mortality
Weekly Report (MMWR) 73:118–123. http://dx.doi.org/10.15585/mmwr.mm7306a1.
Kugeler, K.J., G.M. Farley, J.D. Forrester, and P.S. Mead. 2015. “Geographic Distribution and Expansion
of Human Lyme Disease, United States.” Emerging Infectious Diseases 21(8):1455–1457.
https://doi.org/10.3201/eid2108.141878.
Kundzewicz, Z.W., and I. Pińskwar. 2022. “Are Pluvial and Fluvial Floods on the Rise?” Water
14(17):2612. https://doi.org/10.3390/w14172612.
Lahlali, R., M. Taoussi, S.-E. Laasli, G. Gachara, R. Ezzouggari, Z. Belabess, K. Aberkani, et al. 2024.
“Effects of Climate Change on Plant Pathogens and Host-Pathogen Interactions.” Crop and
Environment 3(3):159–170. https://doi.org/10.1016/j.crope.2024.05.003.
Lal, R., R.F. Follett, and J.M. Kimble. 2003. “Achieving Soil Carbon Sequestration in the United States:
A Challenge to the Policy Makers.” Soil Science 168(12).
https://journals.lww.com/soilsci/fulltext/2003/12000/achieving_soil_carbon_sequestration_in_the_uni
ted.1.aspx
Lan, X., K.W. Thoning, and E.J. Dlugokencky. 2022. “Trends in Globally-Averaged CH4, N2O, and SF6
Determined from NOAA Global Monitoring Laboratory Measurements.” NOAA Global Monitoring
Laboratory. https://doi.org/10.15138/P8XG-AA10.
Lancet Countdown. 2023. “2023 Lancet Countdown on Health and Climate Change Policy Brief for the
United States of America.” Beyeler N.S., P. Knappenberger, J.J. Hess, R.N. Salas. Lancet Countdown
U.S. Policy Brief, London, United Kingdom.
Prepublication Copy
Lancet Countdown. 2024. “2024 Lancet Countdown on Health and Climate Change Policy Brief for the
United States of America.” Beyeler N.S., J. Buonocore, J. Gohlke, S. Johnson, R.N. Salas, J.J. Hess.
Lancet Countdown U.S. Policy Brief, London, United Kingdom.
Langford, A.O., C.J. Senff, R.J. Alvarez II, K.C. Aikin, R. Ahmadov, W.M. Angevine, S. Baidar, et al.
2023. “Were Wildfires Responsible for the Unusually High Surface Ozone in Colorado During
2021?” Journal of Geophysical Research: Atmospheres 128(12).
https://doi.org/10.1029/2022JD037700.
Larson, P.S., A.L. Steiner, M.S. O’Neill, A.P. Baptist, and C.J. Gronlund. 2025. “Chronic and Infectious
Respiratory Mortality and Short-Term Exposures to Four Types of Pollen Taxa in Older Adults in
Michigan, 2006-2017.” BMC Public Health 25(1):173. https://doi.org/10.1186/s12889-025-21386-3
Laufkötter, C., J. Zscheischler, and T.L. Frölicher. 2020. “High-Impact Marine Heatwaves Attributable to
Human-Induced Global Warming.” Science 369(6511):1621–1625.
https://doi.org/10.1126/science.aba0690.
Law, B.E., J.T. Abatzoglou, C.R. Schwalm, D. Byrne, N. Fann, and N.J. Nassikas. 2025. “Anthropogenic
Climate Change Contributes to Wildfire Particulate Matter and Related Mortality in the United
States.” Communications Earth & Environment 6(336). https://doi.org/10.1038/s43247-025-02314-0.
Le Bris, A., K.E. Mills, R.A. Wahle, Y. Chen, M.A. Alexander, A.J. Allyn, J.G. Schuetz, J.D. Scott, and
A.J. Pershing. 2018. “Climate Vulnerability and Resilience in the Most Valuable North American
Fishery.” Proceedings of the National Academy of Sciences of the United States of America
115(8):1831–1836. https://doi.org/10.1073/pnas.1711122115.
L’Ecuyer, T.S., H.K. Beaudoing, M. Rodell, W. Olson, B. Lin, S. Kato, C.A. Clayson, et al. 2015. “The
Observed State of the Energy Budget in the Early Twenty-First Century. Journal of Climate 28:8319–
8346. https://doi.org/10.1175/JCLI-D-14-00556.1.
Lee, S., and H. Ajami. 2023. “Comprehensive Assessment of Baseflow Responses to Long-Term
Meteorological Droughts across the United States.” Journal of Hydrology 626:130256.
https://doi.org/10.1016/j.jhydrol.2023.130256.
Lee, A.S., J. Aguilera, J.A. Efobi, Y.S. Jung, H. Seastedt, M.M. Shah, E. Yang, et al. 2023. “Climate
Change and Public Health: The Effects of Global Warming on the Risk of Allergies and Autoimmune
Diseases.” EMBO Reports 24(4):e56821. https://doi.org/10.15252/embr.202356821
Lee, J., and A.E. Dessler. 2023. “Future Temperature‐Related Deaths in the U.S.: The Impact of Climate
Change, Demographics, and Adaptation.” GeoHealth 7(8):e2023GH000799.
https://doi.org/10.1029/2023GH000799.
Lee, J.-Y., G. Bala, L. Cao, S. Corti, J.P. Dunne, F. Engelbrecht, E. Fischer, et al. 2021. “Future Global
Climate: Scenario-Based Projections and Near-Term Information.” In Climate Change 2021: The
Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the
Intergovernmental Panel on Climate Change. Masson-Delmotte, V., P. Zhai, A. Pirani, S.L. Connors,
C. Péan, S. Berger, N. Caud, Y. Chen, L. Goldfarb, M.I. Gomis, M. Huang, K. Leitzell, E. Lonnoy,
J.B.R. Matthews, T.K. Maycock, T. Waterfield, O. Yelekçi, R. Yu, and B. Zhou, eds. Cambridge
University Press, Cambridge, United Kingdom, and New York, NY, USA: 553–672.
https://doi.org/10.1017/9781009157896.006.
Leeper, R.D., T. Harrington, M.A. Palecki, K. DePolt, E. Scott, J. Runkle, and H.J. Diamond. 2025. “The
Influence of Drought on Heat Wave Intensity, Duration, and Exposure.” Journal of Applied
Meteorology and Climatology 64(5):425–438. https://doi.org/10.1175/JAMC-D-24-0072.1
Lefebvre, K.A., P. Charapata, R. Stimmelmayr, P. Lin, R.S. Pickart, K.A. Hubbard, B.D. Bill, et al. 2025.
“Bowhead Whale Faeces Link Increasing Algal Toxins in the Arctic to Ocean Warming.” Nature
644(8077):693–698. https://doi.org/10.1038/s41586-025-09230-5
Lei, Y., T.-H. Lei, C. Lu, X. Zhang, and F. Wang. 2024. “Wildfire Smoke: Health Effects, Mechanisms,
and Mitigation.” Environmental Science & Technology 58(48):21097–21119.
https://doi.org/10.1021/acs.est.4c06653.
Prepublication Copy
References 91
Leppold, C., L. Gibbs, K. Block, L. Reifels, and P. Quinn. 2022. “Public Health Implications of Multiple
Disaster Exposures.” The Lancet Public Health 7(3):e274–e286. https://doi.org/10.1016/S2468-
2667(21)00255-3.
Lesk, C., W. Anderson, A. Rigden, O. Coast, J. Jägermeyr, S. McDermid, K.F. Davis, and M. Konar.
2022. “Compound Heat and Moisture Extreme Impacts on Global Crop Yields under Climate
Change.” Nature Reviews Earth & Environment 3(12):872–889. https://doi.org/10.1038/s43017-022-
00368-8.
Levy, J., M.K. Woo, S.L. Penn, M. Omary, Y. Tambouret, C.S. Kim and S. Arunachalam. 2016. “Carbon
Reductions and Health Co-Benefits from US Residential Energy Efficiency Measures.”
Environmental Research Letters 11:034017. https://doi.org/10.1088/1748-9326/11/3/034017.
Li, A., H. Luo, Y. Zhu, Z. Zhang, B. Liu, H. Kan, H. Jia, Z. Wu, Y. Guo, and R. Chen. 2025. “Climate
Warming May Undermine Sleep Duration and Quality in Repeated-Measure Study of 23 Million
Records.” Nature Communications 16(1):2609. https://doi.org/10.1038/s41467-025-57781-y
Li, J., M.F. Link, S. Pandit, M.H. Webb, K.J. Mayer, L.A. Garofalo, K.L. Rediger, et al. 2023. “The
Persistence of Smoke VOCs Indoors: Partitioning, Surface Cleaning, and Air Cleaning in a Smoke-
Contaminated House.” Science Advances 9(41):eadh8263.
https://www.science.org/doi/abs/10.1126/sciadv.adh8263.
Li, B., and M. Rodell. 2023. “How Have Hydrological Extremes Changed over the Past 20 Years?”
Journal of Climate 36:8581–8599. https://doi.org/10.1175/JCLI-D-23-0199.1.
Lim, C.C., R.B. Hayes, J. Ahn, Y. Shao, D.T. Silverman, R.R. Jones, C. Garcia, M.L. Bell, and G.D.
Thurston. 2019. “Long-Term Exposure to Ozone and Cause-Specific Mortality Risk in the United
States.” American Journal of Respiratory and Critical Care Medicine 200(8):1022–1031.
https://doi.org/10.1164/rccm.201806-1161OC.
Lin, M., L.W. Horowitz, R. Payton, A.M. Fiore, and G. Tonnesen. 2017. “US Surface Ozone Trends and
Extremes from 1980 to 2014: Quantifying the Roles of Rising Asian Emissions, Domestic Controls,
Wildfires, and Climate.” Atmospheric Chemistry and Physics 17:2943–2970.
https://doi.org/10.5194/acp-17-2943-2017.
Liu, J., B.M. Varghese, A. Hansen, Y. Zhang, T. Driscoll, G. Morgan, K. Dear, M. Gourley, A. Capon,
and P. Bi. 2022a. “Heat Exposure and Cardiovascular Health Outcomes: A Systematic Review and
Meta-Analysis.” The Lancet Planetary Health 6(6):e484–e495. https://doi.org/10.1016/S2542-
5196(22)00117-6.
Liu, J., L.P. Clark, M.J. Bechle, A. Hajat, S.-Y. Kim, A.L. Robinson, L. Sheppard, A.A. Szpiro, and J.D.
Marshall. 2021. “Disparities in Air Pollution Exposure in the United States by Race/Ethnicity and
Income, 1990–2010.” Environmental Health Perspectives 129(12):127005.
https://doi.org/10.1289/EHP8584.
Liu, M., V.R. Patel, and R.K. Wadhera. 2025. “Cold-Related Deaths in the US.” JAMA 333(5):427.
https://doi.org/10.1001/jama.2024.25194.
Loeb, N.G., G.C. Johnson, T.J. Thorsen, J.M. Lyman, F.G. Rose, and S. Kato. 2021. “Satellite and Ocean
Data Reveal Marked Increase in Earth’s Heating Rate.” Geophysical Research Letters
48:e2021GL093047. https://doi.org/10.1029/2021GL093047.
Long, S.P. 1991. “Modification of the Response of Photosynthetic Productivity to Rising Temperature by
Atmospheric CO2 Concentrations: Has its Importance Been Underestimated?” Plant, Cell &
Environment 14(8):729–739. https://doi.org/10.1111/j.1365-3040.1991.tb01439.x
Lwin, K.S., A. Tobias, P.L. Chua, L. Yuan, R. Thawonmas, S. Ith, Z.W. Htay, et al. 2023. “Effects of
Desert Dust and Sandstorms on Human Health: A Scoping Review.” GeoHealth
7(3):e2022GH000728. https://doi.org/10.1029/2022GH000728
Lynch, V.D., J.A. Sullivan, A.B. Flores, X. Xie, S. Aggarwal, R.C. Nethery, M.A. Kioumourtzoglou,
A.E. Nigra, and R.M. Parks. 2025. “Large Floods Drive Changes in Cause-Specific Mortality in the
United States.” Nature Medicine 31(2):663–671. https://doi.org/10.1038/s41591-024-03358-z.
Lynch-Stieglitz, J. 2017. “The Atlantic Meridional Overturning Circulation and Abrupt Climate Change.”
Annual Review of Marine Science 9:83–104. https://doi.org/10.1146/annurev-marine-010816-060415.
Prepublication Copy
Ma, D., L. Gregor, and N. Gruber. 2023. “Four Decades of Trends and Drivers of Global Surface Ocean
Acidification.” Global Biogeochemical Cycles 37:e2023GB007765.
https://doi.org/10.1029/2023GB007765.
Ma, Y., E. Zang, Y. Liu, J. Wei, Y. Lu, H.M. Krumholz, M.L. Bell, and K. Chen. 2023a. “Long-Term
Exposure to Wildland Fire Smoke PM2.5 and Mortality in the Contiguous United States.”
Proceedings of the National Academy of Sciences 121(40):e2403960121.
https://doi.org/10.1101/2023.01.31.23285059
Ma, Y., E. Zang, I. Opara, Y. Lu, H.M. Krumholz, and K. Chen. 2023b. “Racial/Ethnic Disparities in
PM2.5-Attributable Cardiovascular Mortality Burden in the United States.” Nature Human Behaviour
7(12):2074–2083. https://doi.org/10.1038/s41562-023-01694-7.
MacFadden, D.R., S.F. McGough, D. Fisman, M. Santillana, and J.S. Brownstein. 2018. “Antibiotic
Resistance Increases with Local Temperature.” Nature Climate Change 8(6):510–514.
https://doi.org/10.1038/s41558-018-0161-6
Mahendran, R., K. Ju, Z. Yang, Y. Gao, W. Huang, W. Yu, Y. Liu, et al. 2025. “Wildfire-Related Air
Pollution and Infectious Diseases: Systematic Review and Meta-Analysis.” ACS Environmental Au
5(3):253–266. https://doi.org/10.1021/acsenvironau.4c00087.
Mahendran, R., R. Xu, S. Li, and Y. Guo. 2021. “Interpersonal Violence Associated with Hot Weather.”
The Lancet Planetary Health 5(9):e571–e572. https://doi.org/10.1016/S2542-5196(21)00210-2.
Maiga, A.W., S. Deppen, B.K. Scaffidi, J. Baddley, M.C. Aldrich, R.S. Dittus, and E.L. Grogan. 2018.
“Mapping Histoplasma apsulatum Exposure, United States.” Emerging Infectious Diseases
24(10):1835–1839. https://doi.org/10.3201/eid2410.180032.
Makopoulou, E., O. Karjalainen, P. Lipovsky, A. Blais-Stevens, and J. Hjort. 2025. “Susceptibility of
Active-Layer Detachment Failures and Vulnerability of Infrastructure in Alaska and Northwestern
Canada.” Landslides. https://doi.org/10.1007/s10346-025-02603-x.
Makrufardi, F., A. Manullang, D. Rusmawatiningtyas, K.F. Chung, S.C. Lin, and H.C. Chuang. 2023.
“Extreme Weather and Asthma: A Systematic Review and Meta-Analysis.” European Respiratory
Review: An Official Journal of the European Respiratory Society 32(168):230019.
https://doi.org/10.1183/16000617.0019-2023.
Mampage, C.B.A., D.D. Hughes, L.M. Jones, N. Metwali, P.S. Thorne, and E.A. Stone. 2022.
“Characterization of Sub-Pollen Particles in Size-Resolved Atmospheric Aerosol Using Chemical
Tracers.” Atmospheric Environment: X 15:100177. https://doi.org/10.1016/j.aeaoa.2022.100177.
Manabe, S., and R.T. Wetherald. 1967. “Thermal Equilibrium of the Atmosphere with a Given
Distribution of Relative Humidity.” Journal of the Atmospheric Sciences 24:241–
259. https://doi.org/10.1175/1520-0469(1967)024<0241:TEOTAW>2.0.CO;2.
Manion, M., C. Zarakas, S. Wnuck, J. Haskell, A. Belova, D. Cooley, J. Dorn, M. Hoer, and L. Mayo.
2017. “Analysis of the Public Health Impacts of the Regional Greenhouse Gas Initiative.” Abt
Associates 11.
Mansouri, A., W. Wei, J.-M. Alessandrini, C. Mandin, and P. Blondeau. 2022. “Impact of Climate
Change on Indoor Air Quality: A Review.” International Journal of Environmental Research and
Public Health 19(23):15616. https://doi.org/10.3390/ijerph192315616.
Martin, S.S., A.W. Aday, N.B. Allen, Z.I. Almarzooq, C.A.M. Anderson, P. Arora, C.L. Avery, et al.
2025. “2025 Heart Disease and Stroke Statistics: A Report of US and Global Data from the American
Heart Association.” American Heart Journal 151(8):e41–e660.
https://doi.org/10.1161/CIR.0000000000001303.
Martin, J.T., G.T. Pederson, C.A. Woodhouse, E.R. Cook, G.J. McCabe, K.J. Anchukaitis, E.K. Wise, et
al. 2020. “Increased Drought Severity Tracks Warming in the United States’ Largest River Basin.”
Proceedings of the National Academy of Sciences 117(21):11328–11336.
https://doi.org/10.1073/pnas.1916208117.
Martínez-Botí, M.A., G.L. Foster, T.B. Chalk, E.J. Rohling, P.F. Sexton, D.J. Lunt, R.D. Pancost, M.P.S.
Badger, and D.N. Schmidt. 2015. “Plio-Pleistocene Climate Sensitivity Evaluated Using High-
Resolution CO2 Records.” Nature 518(7537):49–54. https://doi.org/10.1038/nature14145.
Prepublication Copy
References 93
Marvel, K., W. Su, R. Delgado, S. Aarons, A. Chatterjee, M.E. Garcia, Z. Hausfather, et al. 2023.
“Chapter 2: Climate Trends.” In Fifth National Climate Assessment, A.R. Crimmins, C.W. Avery,
D.R. Easterling, K.E. Kunkel, B.C. Stewart, and T.K. Maycock, eds. U.S. Global Change Research
Program, Washington, DC, USA. https://doi.org/10.7930/NCA5.2023.CH2.
Marvel, K., B.I. Cook, C. Bonfils, J.E. Smerdon, A.P. Williams, and H. Liu. 2021. “Projected Changes to
Hydroclimate Seasonality in the Continental United States.” Earth’s Future 9(9):e2021EF002019.
https://doi.org/10.1029/2021EF002019.
Mason, R.E., J.M. Craine, N.K. Lany, M. Jonard, S.V. Ollinger, P.M. Groffman, R.W. Fulweiler, et al.
2022. “Evidence, Causes, and Consequences of Declining Nitrogen Availability in Terrestrial
Ecosystems.” Science 376(6590):eabh3767. https://doi.org/10.1126/science.abh3767
Matthes, K., B. Funke, M.E. Andersson, L. Barnard, J. Beer, P. Charbonneau, M.A. Clilverd, et al. 2017.
“Solar Forcing for CMIP6 (v3. 2).” Geoscientific Model Development 10(6):2247–2302.
https://doi.org/10.5194/gmd-10-2247-2017.
Matthews, H.D., and A.J. Weaver. 2010. “Committed Climate Warming.” Nature Geosciences 3:142–
143. https://doi.org/10.1038/ngeo813
Mauritsen, T., Y. Tsushima, B. Meyssignac, N.G. Loeb, M. Hakuba, P. Pilewskie, J. Cole, et al. 2025.
“Earth’s Energy Imbalance More Than Doubled in Recent Decades.” AGU Advances
6:e2024AV001636. https://doi.org/0.1029/2024AV001636.
May, C.L., M.S. Osler, H.F. Stockdon, P.L. Barnard, J.A. Callahan, R.C. Collini, C.M. Ferreira, et al.
2023. “Chapter 9: Coastal Effects.” In Fifth National Climate Assessment. Crimmins, A.R., C.W.
Avery, D.R. Easterling, K.E. Kunkel, B.C. Stewart, and T.K. Maycock, eds. U.S. Global Change
Research Program, Washington, DC, USA. https://doi.org/10.7930/NCA5.2023.CH9.
Maya, S., N. Thakur, T. Benmarhnia, S.D. Weiser, and J.G. Kahn. 2024. “The Impact of Wildfire Smoke
on Asthma Control in California: A Microsimulation Approach.” GeoHealth 8(10):e2024GH001037.
https://doi.org/10.1029/2024GH001037.
McAllister, C., A. Stephens, and S.M. Milrad. 2022. “The Heat Is On: Observations and Trends of Heat
Stress Metrics during Florida Summers.” Journal of Applied Meteorology Climatology 61:277–
296. https://doi.org/10.1175/JAMC-D-21-0113.1.
McArdle, C.E., T.C. Dowling, K. Carey, J. DeVies, D. Johns, A.L. Gates, Z. Stein, et al. 2023. “Asthma-
Associated Emergency Department Visits During the Canadian Wildfire Smoke Episodes — United
States, April– August 2023.” Morbidity and Mortality Weekly Report (MMWR) 72(34):926–932.
https://doi.org/10.15585/mmwr.mm7234a5.
McAvoy, T.J., J. Régnière, R. St-Amant, N.F. Schneeberger, and S.M. Salom. 2017. “Mortality and
Recovery of Hemlock Woolly Adelgid (Adelges tsugae) in Response to Winter Temperatures and
Predictions for the Future.” Forests 8(12):497. https://doi.org/10.3390/f8120497.
McCabe, R.M., B.M. Hickey, R.M. Kudela, K.A. Lefebvre, N.G. Adams, B.D. Bill, F.M.D. Gulland, R.E.
Thomson, W.P. Cochlan, and V.L. Trainer. 2016. “An Unprecedented Coastwide Toxic Algal Bloom
Linked to Anomalous Ocean Conditions.” Geophysical Research Letters 43:10,366–10,376.
https://doi.org/10.1002/2016GL070023.
McCahey, D. 2025. “Historiography of the International Geophysical Year (1957–1958).” In: Handbook
of the Historiography of the Earth and Environmental Sciences. Aronova, E., D. Sepkoski, M.
Tamborini, eds. Historiographies of Science. Springer, Cham. https://doi.org/10.1007/978-3-030-
92679-3_30-1.
McCollum, D.W., J.A. Tanaka, J.A. Morgan, J.E. Mitchell, W.E. Fox, and K.A. Maczko, L. Hidinger,
C.S. Duke, and U.P. Kreuter. 2017. “Climate Change Effects on Rangelands and Rangeland
Management: Affirming the Need for Monitoring.” Ecosystem Health and Sustainability 3(3):e01264.
https://doi.org/10.1002/ehs2.1264.
McDowell, N.G., C.D. Allen, K. Anderson-Teixeira, B.H. Aukema, B. Bond-Lamberty, L. Chini, J.S.
Clark, et al. 2020. “Pervasive Shifts in Forest Dynamics in a Changing World.” Science
368(6494):eaaz9463. https://doi.org/10.1126/science.aaz9463.
Prepublication Copy
McDuffie, E.E., M.C. Sarofim, W. Raich, M. Jackson, H. Roman, K. Seltzer, B.H. Henderson, et al.
2023. “The Social Cost of Ozone‐Related Mortality Impacts from Methane Emissions.” Earth’s
Future 11(9):e2023EF003853. https://doi.org/10.1029/2023EF003853.
McGough, S.F., D.R. MacFadden, M.W. Hattab, K. Mølbak, and M. Santillana. 2020. “Rates of Increase
of Antibiotic Resistance and Ambient Temperature in Europe: A Cross-National Analysis of 28
Countries between 2000 and 2016.” Eurosurveillance 25(45):1900414.
https://www.eurosurveillance.org/content/10.2807/1560-7917.ES.2020.25.45.1900414.
McIntosh, M.M., J.L. Holechek, S.A. Spiegal, A.F. Cibils, and R.E. Estell. 2019. “Long-Term Declining
Trends in Chihuahuan Desert Forage Production in Relation to Precipitation and Ambient
Temperature.” Rangeland Ecology & Management 72(6):976–987.
https://doi.org/10.1016/j.rama.2019.06.002.
Mears, C.A., and F.J. Wentz. 2016. “Sensitivity of Satellite-Derived Tropospheric Temperature Trends to
the Diurnal Cycle Adjustment.” Journal of Climate 29(10):3629–3646. https://doi.org/10.1175/JCLI-
D-15-0744.1.
Meehl, G.A., H. Teng, N. Rosenbloom, A. Hu, C. Tebaldi, and G. Walton. 2022. “How the Great Plains
Dust Bowl Drought Spread Heat Extremes around the Northern Hemisphere.” Scientific Reports
12(1):17380. https://doi.org/10.1038/s41598-022-22262-5.
Metz, M.R., K.M. Frangioso, A.C. Wickland, R.K. Meentemeyer, and D.M. Rizzo. 2012. “An Emergent
Disease Causes Directional Changes in Forest Species Composition in Coastal California.” Ecosphere
3(10):1–23. https://doi.org/10.1890/ES12-00107.1.
Meyssignac, B., T. Boyer, Z. Zhao, M.Z. Hakuba, F.W. Landerer, D. Stammer, A. Köhl, et al. 2019.
“Measuring Global Ocean Heat Content to Estimate the Earth Energy Imbalance.” Frontiers in
Marine Science 6:432. https://doi.org/10.3389/fmars.2019.00432.
Milly, P.C.D., and K.A. Dunne. 2020. “Colorado River Flow Dwindles as Warming-Driven Loss of
Reflective Snow Energizes Evaporation.” Science 367(6483):1252–1255.
https://doi.org/10.1126/science.aay9187.
Mishra, V., and K.A. Cherkauer. 2010. “Retrospective Droughts in the Crop Growing Season:
Implications to Corn and Soybean Yield in the Midwestern United States.” Agricultural and Forest
Meteorology 150(7–8):1030–1045. https://doi.org/10.1016/j.agrformet.2010.04.002.
Molaei, G., E.A.H. Little, S.C. Williams, and K.C. Stafford. 2019. “Bracing for the Worst — Range
Expansion of the Lone Star Tick in the Northeastern United States.” New England Journal of
Medicine 381(23):2189–2192. https://doi.org/10.1056/NEJMp1911661.
Montillet, J.P., W. Finsterle, G. Kermarrec, R. Sikonja, M. Haberreiter, W. Schmutz, and T. Dudok de
Wit. 2022. “Data Fusion of Total Solar Irradiance Composite Time Series Using 41 Years of Satellite
Measurements.” Journal of Geophysical Research: Atmospheres 127(13):e2021JD036146.
https://doi.org/10.1002/essoar.10508721.2
Montrose, L., A. Schuller, S. D’Evelyn, and C. Migliaccio. 2022. “Chapter 11: Wildfire Smoke
Toxicology and Health.” In Landscape Fire, Smoke, and Health: Linking Biomass Burning Emissions
to Human Well-Being, Loboda, T.V., N.H.F. French, and R.C. Puett, eds.
https://doi.org/10.1002/essoar.10512251.1.
Moore, F.C., U. Baldos, T. Hertel, and D. Diaz. 2017. “New Science of Climate Change Impacts on
Agriculture Implies Higher Social Cost of Carbon.” Nature Communications 8(1):1607.
https://doi.org/10.1038/s41467-017-01792-x.
Mordecai, E.A., J.M. Caldwell, M.K. Grossman, C.A. Lippi, L.R. Johnson, M. Neira, J.R. Rohr, et al.
2019. “Thermal Biology of Mosquito‐Borne Disease.” Ecology Letters 22(10):1690–1708.
https://doi.org/10.1111/ele.13335.
Morford, S.L., B.W. Allred, D. Twidwell, M.O. Jones, J.D. Maestas, C.P. Roberts, and D.E. Naugle.
2022. "Herbaceous Production Lost to Tree Encroachment in United States Rangelands." Journal of
Applied Ecology 59:2971–2982. https://doi.org/10.1111/1365-2664.14288.
Mulliken, J.S., K.N. Hampshire, A.G. Rappold, M. Fung, J.M. Babik, and S.B. Doernberg. 2023. “Risk of
Systemic Fungal Infections after Exposure to Wildfires: A Population-Based, Retrospective Study in
Prepublication Copy
References 95
Prepublication Copy
NASEM. 2016. Attribution of Extreme Weather Events in the Context of Climate Change. Washington,
DC: The National Academies Press. https://doi.org/10.17226/21852.
National Oceanic and Atmospheric Administration (NOAA). 2025. “Climate Change: Ocean Heat
Content.” https://www.climate.gov/news-features/understanding-climate/climate-change-ocean-heat-
content
NOAA. 2024. “What percentage of the American population lives near the coast?” Retrieved September
3, 2025, from https://oceanservice.noaa.gov/facts/population.html.
NOAA. 2024a. “During a Year of Extremes, Carbon Dioxide Levels Surge Faster Than Ever: The Two-
Year Increase in Keeling Curve Peak is the Largest on Record.” June 6. https://www.noaa.gov/news-
release/during-year-of-extremes-carbon-dioxide-levels-surge-faster-than-ever.
NOAA. 2024b. NOAA Monthly U.S. Climate Divisional Database
(nClimDiv). www1.ncdc.noaa.gov/pub/data/cirs/climdiv.
NOAA. 2023. “Climate Change: Global Sea Level.” NOAA Climate.gov. August 22.
https://www.climate.gov/news-features/understanding-climate/climate-change-global-sea-level.
National Petroleum Council. 2024. “Chapter 3: Detection and Quantitative Tracking of Greenhouse Gas
Emissions from Oil and Gas Supply Chains.” In Charting the Course.
https://chartingthecourse.npc.org/files/GHG-V2_Chapter_3-FINAL.pdf.
National Research Council (NRC). 2020. “Climate Change: Evidence and Causes: Update 2020.”
Washington, DC: The National Academies Press. https://doi.org/10.17226/25733.
NRC. 2013. Abrupt Impacts of Climate Change: Anticipating Surprises. Washington, DC: The National
Academies Press. https://doi.org/10.17226/18373.
NRC. 1979. Carbon Dioxide and Climate: A Scientific Assessment. Washington, DC: National Academy
Press. https://doi.org/10.17226/19856.
Newbury, J.B., L. Arseneault, S. Beevers, N. Kitwiroon, S. Roberts, C.M. Pariante, F.J. Kelly, and H.L.
Fisher. 2019. “Association of Air Pollution Exposure with Psychotic Experiences During
Adolescence.” JAMA Psychiatry 76(6):614. https://doi.org/10.1001/jamapsychiatry.2019.0056.
Norby, R.J., E.H. DeLucia, B. Gielen, C. Calfapietra, C.P. Giardina, J.S. King, J. Ledford, et al. 2005.
“Forest Response to Elevated CO2 is Conserved Across a Broad Range of Productivity.” Proceedings
of the National Academy of Sciences 102(50):18052–18056.
https://doi.org/10.1073/pnas.0509478102.
Nordio, F., A. Zanobetti, E. Colicino, I. Kloog, and J. Schwartz. 2015. “Changing Patterns of the
Temperature-Mortality Association by Time and Location in the US, and Implications for Climate
Change.” Environment International 81:80–86. https://doi.org/10.1016/j.envint.2015.04.009
Nori-Sarma, A., S. Sun, Y. Sun, K.R. Spangler, R. Oblath, S. Galea, J.L. Gradus, and G.A. Wellenius.
2022. “Association Between Ambient Heat and Risk of Emergency Department Visits for Mental
Health Among US Adults, 2010 to 2019.” JAMA Psychiatry 79(4):341.
https://doi.org/10.1001/jamapsychiatry.2021.4369.
Okamoto-Mizuno, K., and K. Mizuno. 2012. “Effects of Thermal Environment on Sleep and Circadian
Rhythm.” Journal of Physiological Anthropology 31(1):14. https://doi.org/10.1186/1880-6805-31-14.
O’Neill, B.C., C. Tebaldi, D.P. van Vuuren, V. Eyring, P. Friedlingstein, G. Hurtt, R. Knutti, et al. 2016.
“The Scenario Model Intercomparison Project (ScenarioMIP) for CMIP6.” Geoscientific Model
Development 9:3461–3482. https://doi.org/10.5194/gmd-9-3461-2016.
Orr, A., N.B. Alden, E. Austin, Z. Jaffar, J. Knudson, J. Graham, C.T. Migliaccio, et al. 2025. “Wildfire-
Season Fine Particulate Matter Exposure and Associations with Influenza and Influenza-Like-Illness
Risk in the Western USA.” Environmental Health Perspectives.
https://ehp.niehs.nih.gov/doi/abs/10.1289/EHP16607.
Ostoja, S.M., A.R. Crimmins, R.G. Byron, A.E. East, M. Méndez, S.M. O’Neill, D.L. Peterson, et al.
2023. “Focus on Western Wildfires.” In Fifth National Climate Assessment, Crimmins, A.R., C.W.
Avery, D.R. Easterling, K.E. Kunkel, B.C. Stewart, and T.K. Maycock, eds. U.S. Global Change
Research Program. https://doi.org/10.7930/NCA5.2023.F2.
Prepublication Copy
References 97
Overpeck, J.T., and B. Udall. 2020. “Climate Change and the Aridification of North America.”
Proceedings of the National Academy of Sciences 117(22):11856–11858.
https://doi.org/10.1073/pnas.2006323117.
Pan, Y., R.A. Birdsey, J. Fang, R. Houghton, P.E. Kauppi, W.A. Kurz, O.L. Phillips, et al. 2011. “A
Large and Persistent Carbon Sink in the World’s Forests.” Science 333(6045):988–993.
https://doi.org/10.1126/science.1201609.
Parsons, L.A., F. Lo, A. Ward, D. Shindell, and S.R. Raman. 2023. “Higher Temperatures in Socially
Vulnerable US Communities Increasingly Limit Safe Use of Electric Fans for Cooling.” GeoHealth
7(8):e2023GH000809. https://doi.org/10.1029/2023GH000809.
Parsons, L.A., D. Shindell, M. Tigchelaar, Y. Zhang, and J.T. Spector. 2021. “Increased Labor Losses and
Decreased Adaptation Potential in a Warmer World.” Nature Communications 12(1):7286.
https://doi.org/10.1038/s41467-021-27328-y.
Paudel, B., T. Chu, M. Chen, V. Sampath, M. Prunicki, and K.C. Nadeau. 2021. “Increased Duration of
Pollen and Mold Exposure are Linked to Climate Change.” Scientific Reports 11(1):12816.
https://doi.org/10.1038/s41598-021-92178-z.
Pavlovic, N.R., S.Y. Chang, K.J. Craig, C.R. Scarborough, J.G. Coughlin, J.D. Herrick, and C.T. Driscoll.
2025. “Quantification of Ozone Exposure Impacts and Their Uncertainties on Growth and Survival of
88 Tree Species Across the United States.” Journal of Geophysical Research: Atmospheres
130(12):e2024JD042063. https://doi.org/10.1029/2024JD042063.
Pendall, E., D. Bachelet, R. Conant, B. El Masri, L. Flanagan, A. Knapp, J. Liu, J., S. Liu, and S.
Schaeffer. 2018. “Chapter 10: Grasslands.” In Second State of the Carbon Cycle Report (SOCCR2): A
Sustained Assessment Report. Cavallaro, N., G. Shrestha, R. Birdsey, M.A. Mayes, R.G. Najjar, S.C.
Reed, P. Romero-Lankao, and Z. Zhu, eds. 399–427. U.S. Global Change Research Program.
https://doi.org/10.7930/SOCCR2.2018.Ch10.
Peralta, A.A., M.S. Link, J. Schwartz, H. Luttmann-Gibson, D.W. Dockery, A. Blomberg, Y. Wei, et al.
2020. “Exposure to Air Pollution and Particle Radioactivity with the Risk of Ventricular
Arrhythmias.” Circulation 142(9):858–867.
https://doi.org/10.1161/CIRCULATIONAHA.120.046321.
Pershing, A., K. Mills, A. Dayton, B. Franklin, and B. Kennedy. 2018. “Evidence for Adaptation from the
2016 Marine Heatwave in the Northwest Atlantic Ocean.” Oceanography 31(2):152–161.
https://doi.org/10.5670/oceanog.2018.213.
Plaas, H.E., and H.W. Paerl. 2021. “Toxic Cyanobacteria: A Growing Threat to Water and Air Quality.”
Environmental Science & Technology 55(1):44–64. https://doi.org/10.1021/acs.est.0c06653
Polley, H.W., D.D. Briske, J.A. Morgan, K. Wolter, D.W. Bailey, and J.R. Brown. 2013. “Climate
Change and North American Rangelands: Trends, Projections, and Implications.” Rangeland Ecology
& Management 66:493–511. https://doi.org/10.2111/REM-D-12-00068.1.
Possinger, A.R., C.T. Driscoll, M.B. Green, T.J. Fahey, C.E. Johnson, M.M.K. Koppers, L.D. Martel, et
al. 2025. “Increasing Soil Respiration in a Northern Hardwood Forest Indicates Symptoms of a
Changing Carbon Cycle.” Communications Earth & Environment 6(1):418.
https://doi.org/10.1038/s43247-025-02405-y.
Preston, D.L., J.A. Mischler, A.R. Townsend, and P.T.J. Johnson. 2016. “Disease Ecology Meets
Ecosystem Science.” Ecosystems 19(4):737–748. https://doi.org/10.1007/s10021-016-9965-2.
Purich, A., and E.W. Doddridge. 2023. “Record Low Antarctic Sea Ice Coverage Indicates a New Sea Ice
State.” Communications Earth & Environment 4(314). https://doi.org/10.1038/s43247-023-00961-9.
Putman, A.L., M. Blakowski, D. DiViesti, D. Fernandez, M. McDonnell, P. Longley, and D.K. Jones.
2025. “Contributions of Great Salt Lake Playa‐ and Industrially Sourced Priority Pollutant Metals in
Dust Contribute to Possible Health Hazards in the Communities of Northern Utah.” GeoHealth
9(8):e2025GH001462. https://doi.org/10.1029/2025GH001462.
Qian, Y., and T. Liu. 2025. “Heat Vulnerability Assessment: A Systematic Review of Critical Metrics.”
Hygiene and Environmental Health Advances 15:100138.
https://doi.org/10.1016/j.heha.2025.100138.
Prepublication Copy
Raabe, E.A., and R.P. Stumpf. 2016. “Expansion of Tidal Marsh in Response to Sea-Level Rise: Gulf
Coast of Florida, USA.” Estuaries and Coasts 39(1):145–157. https://doi.org/10.1007/s12237-015-
9974-y.
Rahman, M.M., R. McConnell, H. Schlaerth, J. Ko, S. Silva, F.W. Lurmann, L. Palinkas, et al. 2022.
“The Effects of Coexposure to Extremes of Heat and Particulate Air Pollution on Mortality in
California: Implications for Climate Change.” American Journal of Respiratory and Critical Care
Medicine 206(9):1117–1127. https://doi.org/10.1164/rccm.202204-0657OC.
Rai, M., M. Stafoggia, F. de’Donato, M. Scortichini, S. Zafeiratou, L. Vazquez Fernandez, S. Zhang, et
al. 2023. “Heat-Related Cardiorespiratory Mortality: Effect Modification by Air Pollution across 482
Cities from 24 Countries.” Environment International 174:107825.
https://doi.org/10.1016/j.envint.2023.107825.
Rajagopalan, S., and P.J. Landrigan. 2021. “Pollution and the Heart.” New England Journal of Medicine
385(20):1881–1892. https://doi.org/10.1056/NEJMra2030281.
Rao, C.Y., M.A. Riggs, G.L. Chew, M.L. Muilenberg, P.S. Thorne, D. Van Sickle, K.H. Dunn, and C.
Brown. 2007. “Characterization of Airborne Molds, Endotoxins, and Glucans in Homes in New
Orleans after Hurricanes Katrina and Rita.” Applied and Environmental Microbiology 73(5):1630–
1634. https://doi.org/10.1128/AEM.01973-06.
Rappazzo, K.M., J.L. Nichols, R.B. Rice, and T.J. Luben. 2021. “Ozone Exposure During Early
Pregnancy and Preterm Birth: A Systematic Review and Meta-Analysis.” Environmental Research
198:111317. https://doi.org/10.1016/j.envres.2021.111317.
Rau, A., A.K. Baldomero, J.E. Bell, J. Rennie, C.H. Wendt, G.A.M. Tarr, B.H. Alexander, and J.D.
Berman. 2025. “Compound Drought and Heatwave Extreme Weather Events: Mortality Risk in
Individuals with Chronic Respiratory Disease.” Environmental Epidemiology 9(3):e389.
https://doi.org/10.1097/EE9.0000000000000389
Reeves, M.C., B.B. Hanberry, H. Wilmer, N.E. Kaplan, and W.K. Lauenroth. 2021 “An Assessment of
Production Trends on the Great Plains from 1984 to 2017.” Rangeland Ecology & Management
78:165–179. https://doi.org/10.1016/j.rama.2020.01.011.
Remigio, R.V., H. He, J. Raimann, P. Kotanko, F.W. Maddux, A.R. Sapkota, X.-Z. Liang, R. Puett, X.
He, and A. Sapkota. 2021. “Combined Effects of Air Pollution and Extreme Heat Events Among
ESKD Patients Within the Northeastern United States.” Science of the Total Environment
812:152481. https://doi.org/10.1016/j.scitotenv.2021.152481.
Riahi, K., D.P. van Vuuren, E. Kriegler, J. Edmonds, B.C. O’Neill, S. Fujimori, N. Bauer, et al. 2017.
“The Shared Socioeconomic Pathways and Their Energy, Land Use, and Greenhouse Gas Emissions
Implications: An Overview.” Global Environmental Change 42:153–168.
https://doi.org/10.1016/j.gloenvcha.2016.05.009.
Rignot, E., J. Mouginot, B. Scheuchl, M. van den Broeke, M.J. van Wessem, and M. Morlighem. 2019.
“Four Decades of Antarctic Ice Sheet Mass Balance from 1979–2017.” Proceedings of the National
Academy of Sciences 116(4):1095–1103. https://doi.org/10.1073/pnas.1812883116.
Rochlin, I., D.V. Ninivaggi, M.L. Hutchinson, and A. Farajollahi. 2013. “Climate Change and Range
Expansion of the Asian Tiger Mosquito (Aedes albopictus) in Northeastern USA: Implications for
Public Health Practitioners.” PLoS ONE 8(4):e60874. https://doi.org/10.1371/journal.pone.0060874.
Rode, A., T. Carleton, M. Delgado, M. Greenstone, T. Houser, S. Hsiang, A. Hultgren, et al. 2021.
“Estimating a Social Cost of Carbon for Global Energy Consumption.” Nature 598(7880):308–314.
https://doi.org/10.1038/s41586-021-03883-8.
Rodell, M., H.K. Beaudoing, T.S. L’Ecuyer, W.S. Olson, J.S. Famiglietti, P.R. Houser, R. Adler, et al.
2015. “The Observed State of the Water Cycle in the Early Twenty-First Century.” Journal of
Climate 28:8289–8318. https://doi.org/10.1175/JCLI-D-14-00555.1.
Romanello, M., M. Walawender, S.-C. Hsu, A. Moskeland, Y. Palmeiro-Silva, D. Scamman, Z. Ali, et al.
2024. “The 2024 Report of the Lancet Countdown on Health and Climate Change: Facing Record-
Breaking Threats from Delayed Action.” The Lancet 404(10465):1847–1896.
https://doi.org/10.1016/S0140-6736(24)01822-1.
Prepublication Copy
References 99
Prepublication Copy
100 Effects of Human-Caused GHG Emissions on U.S. Climate, Health, and Welfare
Schets, F.M., I.E. Pol-Hofstad, H.H.J.L. van den Berg, and J.F. Schijven. 2025. “Climate Change-Related
Temperature Impact on Human Health Risks of Vibrio Species in Bathing and Surface Water.”
Microorganisms 13(8). https://doi.org/10.3390/microorganisms13081893.
Schieder, N.W., D.C. Walters, and M.L. Kirwan. 2018. “Massive Upland to Wetland Conversion
Compensated for Historical Marsh Loss in Chesapeake Bay, USA.” Estuaries and Coasts 41:940–
951. https://doi.org/10.1007/s12237-017-0336-9.
Schlenker, W., and M.J. Roberts. 2009. “Nonlinear Temperature Effects Indicate Severe Damages to U.S.
Crop Yields Under Climate Change.” Proceedings of the National Academy of Sciences
106(37):15594–15598. https//doi.org/10.1073/pnas.0906865106.
Schreck, C. 2025. “Sidebar 4.1: Hurricane Helene: Inside Western North Carolina’s Historic Flood. [In
State of the Climate in 2024]” Bulletin in the American Meteorological Society 106(8):S284–S286.
https://doi.org/10.1175/BAMS-D-25-0086.1
Scussolini, P., L.N. Luu, S. Philip, W.R. Berghuijs, D. Eilander, J.C.J.H. Aerts, S.F. Kew, et al. 2024.
“Challenges in the Attribution of River Flood Events.” WIREs Climate Change 15(3):e874.
https://doi.org/10.1002/wcc.874.
Sergi, B.J., P.J. Adams, N.Z. Muller, A.L. Robinson, S.J. Davis, J.D. Marshall, and I.L. Azevedo. 2020.
“Optimizing Emissions Reductions from the U.S. Power Sector for Climate and Health Benefits.”
Environmental Science & Technology 54(12):7513–7523. https://doi.org/10.1021/acs.est.9b06936
Shah, A.S., J.P. Langrish, H. Nair, D.A. McAllister, A.L. Hunter, K. Donaldson, D.E. Newby, and N.L.
Mills. 2013. “Global Association of Air Pollution and Heart Failure: A Systematic Review and Meta-
Analysis.” The Lancet 382(9897):1039–1048. https://doi.org/10.1016/S0140-6736(13)60898-3.
Shi, J.-R., B.D. Santer, Y.-O. Kown, and S. Wijffels. 2024. “The Emerging Human Influence on the
Seasonal Cycle of Seas Surface Temperature.” Nature Climate Change 14(4):1–9.
https://doi.org/10.1038/s41558-024-01958-8.
Shindell, D., P. Sadavarte, I. Aben, T. de Oliveira Bredariol, G. Dreyfus, L. Höglund-Isaksson, B.
Poulter, et al. 2024. “The Methane Imperative.” Frontier Scientific Publishing 2.
https://doi.org/10.3389/fsci.2024.1349770.
Shindell, D., Y. Zhang, M. Scott, M. Ru, K. Stark, and K.L. Ebi. 2020. “The Effects of Heat Exposure on
Human Mortality Throughout the United States.” GeoHealth 4(4):e2019GH000234.
https://doi.org/10.1029/2019GH000234.
Sias, J.E., E.V. Dave, B.S. Underwood, B.F. Bowers, J.T. Harvey, T.F.P. Henning, S.L. Tighe, et al.
2025. “Climate Change Impacts on Roadways.” Nature Reviews Earth & Environment.
https://doi.org/10.1038/s43017-025-00711-9.
Simler-Williamson, A.B., D.M. Rizzo, and R.C. Cobb. 2019. “Interacting Effects of Global Change on
Forest Pest and Pathogen Dynamics.” Annual Review of Ecology, Evolution, and Systematics
50(1):381–403. https://doi.org/10.1146/annurev-ecolsys-110218-024934.
Simpson, I.R., T.A. Shaw, P. Ceppi, A.C. Clement, E. Fischer, K.M. Grise, A.G. Pendergrass, et al. 2025.
“Confronting Earth System Model Trends with Observations.” Science Advances 11(11).
https://doi.org/10.1126/sciadv.adt8035.
Singh, N.K., S.M. Saia, R. Bhattacharya, H. Ajami, and D.M. Borrok. 2025. “Unraveling the Causal
Influences of Drought and Crop Production on Groundwater Levels Across the Contiguous United
States.” PNAS Nexus 4(5):pgaf129. https://doi.org/10.1093/pnasnexus/pgaf129.
Singh, N., A.T. Areal, S. Breitner, S. Zhang, S. Agewall, T. Schikowski, and A. Schneider. 2024. “Heat
and Cardiovascular Mortality: An Epidemiological Perspective.” Circulation Research 134(9):1098–
1112. https://doi.org/10.1161/CIRCRESAHA.123.323615.
Skendžić, S., M. Zovko, I.P. Živković, V. Lešić, and D. Lemić. 2021. “The Impact of Climate Change on
Agricultural Insect Pests.” Insects 12(5):440. https://doi.org/10.3390/insects12050440.
Smale, D.A., T. Wernberg, E.C.J. Oliver, M. Thomsen, B.P. Harvey, S.C. Straub, M.T. Burrows, et al.
2019. “Marine Heatwaves Threaten Global Biodiversity and the Provision of Ecosystem
Services.” Nature Climate Change 9:306–312. https://doi.org/10.1038/s41558-019-0412-1.
Prepublication Copy
References 101
Smith, K.E., M. Aubin, M.T. Burrows, K. Filbee-Dexter, A.J. Hobday, N.J. Holbrook, N.G. King. et al.
2024. “Global Impacts of Marine Heatwaves on Coastal Foundation Species.” Nature
Communications 15:5052. https://doi.org/10.1038/s41467-024-49307-9.
Smith, E.T., and S.C. Sheridan. 2020. “Where Do Cold Air Outbreaks Occur, and How Have They
Changed Over Time?” Geophysical Research Letters 47:e2020GL086983.
https://doi.org/10.1029/2020GL086983.
Smith, M.D. 2011. “The Ecological Role of Climate Extremes: Current Understanding and Future
Prospects.” Journal of Ecology 99(3):651–655. https://doi.org/10.1111/j.1365-2745.2011.01833.x.
Song, J., S. Wan, S. Piao, A.K. Knapp, A.T. Classen, S. Vicca, P. Ciais, et al. 2019. “A Meta-Analysis of
1,119 Manipulative Experiments on Terrestrial Carbon-Cycling Responses to Global Change.”
Nature Ecology & Evolution 3(9):1309–1320. https://doi.org/10.1038/s41559-019-0958-3.
Spies, I., K.M. Gruenthal, D.P. Drinan, A.B. Hollowed, D.E. Stevenson, C.M. Tarpey, and L. Hauser.
2020. “Genetic Evidence of a Northward Range Expansion in the Eastern Bering Sea Stock of Pacific
Cod.” Evolutionary Applications 13(2):362–375. https://doi.org/10.1111/eva.12874.
Stafoggia, M., P. Michelozzi, A. Schneider, B. Armstrong, M. Scortichini, M. Rai, S. Achilleos, et al.
2023. “Joint Effect of Heat and Air Pollution on Mortality in 620 Cities of 36 Countries.”
Environment International 181:108258. https://doi.org/10.1016/j.envint.2023.108258
Stephens, G.L., M.Z. Hakuba, S. Kato, A. Gettelman, J.-L. Dufresne, T. Andrews, J.N.S. Cole, U. Willen,
and T. Mauritsen. 2022. “The Changing Nature of Earth’s Reflected Sunlight.” Proceedings of the
Royal Society A: Mathematical, Physical & Engineering Sciences 478(2263):20220053.
https://doi.org/10.1098/rspa.2022.0053.
Stephens, G.L., M.Z. Hakuba, M.J. Webb, M. Lebsock, Q. Yue, B.H. Kahn, S. Hristova-Veleva, et al.
2018. “Regional Intensification of the Tropical Hydrological Cycle During ENSO.” Geophysical
Research Letters 45(9):4361–4370. https://doi.org/10.1029/2018GL077598.
Stevens, J.P., and Supreme Court Of The United States. U.S. Reports: Massachusetts v. EPA, 549 U.S.
497. 2006. https://supreme.justia.com/cases/federal/us/549/497.
Su, L., Q. Cao, S. Shukla, M. Pan, and D.P. Lettenmaier. 2023. “Evaluation of Subseasonal Drought
Forecast Skill over the Coastal Western United States.” Journal of Hydrometeorology 24(4):709-
726. https://doi.org/10.1175/JHM-D-22-0103.1.
Su, L., Q. Cao, M. Xiao, D.M. Mocko, M. Barlage, D. Li, C.D. Peters-Lidard, and D.P. Lettenmaier.
2021. “Drought Variability over the Conterminous United States for the Past Century. Journal of
Hydrometeorology 22(5):1153–1168. https://doi.org/10.1175/JHM-D-20-0158.1.
Sun, S., K.R. Weinberger, A. Nori-Sarma, K.R. Spangler, Y. Sun, F. Dominici, and G.A. Wellenius.
2021. “Ambient Heat and Risks of Emergency Department Visits Among Adults in the United States:
Time Stratified Case Crossover Study.” BMJ e065653. https://doi.org/10.1136/bmj-2021-065653.
Sun, Q., F. Zwiers, X. Zhang, and J. Yan. 2022 “Quantifying the Human Influence on the Intensity of
Extreme 1- and 5-Day Precipitation Amounts at Global, Continental, and Regional Scales.” Journal
of Climate 35:195–210. https://doi.org/10.1175/JCLI-D-21-0028.1.
Supran, G., S. Rahmstorf. and N. Oreskes, 2023. “Assessing ExxonMobil’s Global Warming
Projections.” Science 379(6628):eabk0063. https://www.science.org/doi/10.1126/science.abk0063
Swain, D.L., A.F. Prein, J.T. Abatzoglou, C.M. Albano, M. Brunner, N.S. Diffenbaugh, D. Singh, C.B.
Skinner, and D. Touma. 2025. “Hydroclimate Volatility on a Warming Earth.” Nature Reviews
Earth & Environment 6:35–50. https://doi.org/10.1038/s43017-024-00624-z.
Sweet, W.V., B.D. Hamlington, R.E. Kopp, C.P. Weaver, P.L. Barnard, D. Bekaert, W. Brooks, et al.
2022. “Global and Regional Sea Level Rise Scenarios for the United States: Updated Mean
Projections and Extreme Weather Level Probabilities along U.S. Coastlines.” NOAA Technical
Report NOS 01. National Oceanic and Atmospheric Administration. 111.
Teixeira, J., R.C. Wilson, and H.T. Thrastarson, 2024. “Direct Observational Evidence from Space of the
Effect of CO2 Increase on Longwave Spectral Radiances: The Unique Role of High-Spectral-
Resolution Measurements.” Atmospheric Chemistry and Physics 24(10):6375–6383.
https://doi.org/10.5194/egusphere-2023-924.
Prepublication Copy
102 Effects of Human-Caused GHG Emissions on U.S. Climate, Health, and Welfare
Thiault, L., C. Mora, J.E. Cinner, W.W.L. Cheung, N.A.J. Graham, F.A. Januchowski-Hartley, D.
Mouillot, U.R. Sumaila, and J. Claudet. 2019 “Escaping the Perfect Storm of Simultaneous Climate
Change Impacts on Agriculture and Marine Fisheries.” Science Advances 5(11):9976.
https://doi.org/10.1126/sciadv.aaw9976.
Thompson, R., E.L. Lawrance, L.F. Roberts, K. Grailey, H. Ashrafian, H. Maheswaran, M.B. Toledano,
and A. Darzi. 2023. “Ambient Temperature and Mental Health: A Systematic Review and Meta-
Analysis.” The Lancet Planetary Health 7(7):e580–e589. https://doi.org/10.1016/S2542-
5196(23)00104-3.
Tong, D. Q., J.X.L. Wang, T.E. Gill, H. Lei, and B. Wang. 2017. “Intensified Dust Storm Activity and
Valley Fever Infection in the Southwestern United States.” Geophysical Research Letters
44(9):4304–4312. https://doi.org/10.1002/2017GL073524.
Townhill, B.L., J. Tinker, M. Jones, S. Pitois, V. Creach, S.D. Simpson, S. Dye, E. Bear, and J.K.
Pinnegar. 2018. “Harmful Algal Blooms and Climate Change: Exploring Future Distribution
Changes.” ICES Journal of Marine Science 75(6):1882–1893. https://doi.org/10.1093/icesjms/fsy113.
Trainer, V.L., S.K. Moore, G. Hallegraeff, R.M. Kudela, A. Clement, J.I. Mardones, and W.P. Cochlan.
2020. “Pelagic Harmful Algal Blooms and Climate Change: Lessons from Nature’s Experiments with
Extremes.” Harmful Algae 91:101591. https://doi.org/10.1016/j.hal.2019.03.009.
Transportation Research Board (TRB). 2014. Strategic Issues Facing Transportation, Volume 2: Climate
Change, Extreme Weather Events, and the Highway System: Practitioner’s Guide and Research
Report. The National Academies Press. https://doi.org/10.17226/22473.
Tripathy, K.P., S. Mukherjee, A.K. Mishra, M.E. Mann, and A.P. Williams. 2023. “Climate Change Will
Accelerate the High-End Risk of Compound Drought and Heatwave Events.” Proceedings of the
National Academy of Sciences 120(28):e2219825120. https://doi.org/10.1073/pnas.2219825120.
Turner, M.C., M. Jerrett, C.A. Pope, D. Krewski, S.M. Gapstur, W.R. Diver, B.S. Beckerman, et al. 2016.
“Long-Term Ozone Exposure and Mortality in a Large Prospective Study.” American Journal of
Respiratory and Critical Care Medicine 193(10):1134–1142. https://doi.org/10.1164/rccm.201508-
1633OC.
Turnock, S.T., D. Akritidis, L. Horowitz, M. Mertens, A. Pozzer, C.L. Reddington, H. Wang, P. Zhou,
and F. O’Connor. 2025. “Drivers of Change in Peak-Season Surface Ozone Concentrations and
Impacts on Human Health over the Historical Period (1850-2014).” Atmospheric Chemistry and
Physics 25:7111–7136.https://doi.org/10.5194/acp-25-7111-2025.
Tyndall, J. 1863. “On Radiation through the Earth’s Atmosphere.” Philosophical Magazine 4(25):200–
206. https://ajsonline.org/api/v1/articles/64993-i-proceedings-of-learned-societies-on-the-radiation-
through-the-earth-s-atmosphere.pdf.
Ummenhofer, C.C., and G.A. Meehl. 2017. “Extreme Weather and Climate Events with Ecological
Relevance: A Review.” Philosophical Transactions of the Royal Society B: Biological Sciences
372(1723):20160135. https://doi.org/10.1098/rstb.2016.0135.
United Nations Environment Programme (UNEP) and Climate and Clean Air Coalition (CCAC). 2021.
“Global Methane Assessment: Benefits and Costs of Mitigating Methane Emissions.”
https://www.unep.org/resources/report/global-methane-assessment-benefits-and-costs-mitigating-
methane-emissions.
UNEP. 2024. “Emissions Gap Report 2024: No More Hot Air… Please! With a Massive Gap Between
Rhetoric and Reality, Countries Draft New Climate Commitments.”
https://doi.org/10.59117/20.500.11822/46404.
U.S. Bureau of Transportation Statistics. 2025. “Vehicle Miles Traveled (VMT).” Retrieved from FRED,
Federal Reserve Bank of St. Louis. https://fred.stlouisfed.org/series/VMT.
U.S. Energy Information Administration (EIA). 2025. “U.S. Electricity Peak Demand Set New Records
Twice in July.” U.S. Energy Information Administration.
https://www.eia.gov/todayinenergy/detail.php?id=65864.
EIA. 2025a. “Table 1.2 Primary Energy Production by Source.” U.S. Energy Information Administration.
https://www.eia.gov/totalenergy/data/browser/?tbl=T01.02#/?f=M.
Prepublication Copy
References 103
EIA. 2025b. “Table 1.3 Primary Energy Consumption by Source.” U.S. Energy Information
Administration. https://www.eia.gov/totalenergy/data/browser/?tbl=T01.02#/?f=M.
U.S. Global Change Research Program (USGCRP). 2023. “Fifth National Climate Assessment.”
Crimmins, A.R. C.W. Avery, D.R. Easterling, K.E. Kunkel, B.C. Stewart and T.K. Maycock, eds. US
Global Change Research Program Washington, DC, USA. https://doi.org/10.7930/NCA5.2023.
USGCRP. 2016. The Impacts of Climate Change on Human Health in the United States: A Scientific
Assessment. Crimmins, A., J. Balbus, J.L. Gamble, C.B. Beard, J.E. Bell, D. Dodgen, R.J. Eisen, N.
Fann, M.D. Hawkins, S.C. Herring, L. Jantarasami, D.M. Mills, S. Saha, M.C. Sarofim, J. Trtanj, and
L. Ziska, eds. U.S. Global Change Research Program: Washington, DC.
Van Eldijk, T.J.B., E.A. Sheridan, G. Martin, F.J. Weissing, O.P. Kuipers, and G.S. Van Doorn. 2024.
“Temperature Dependence of the Mutation Rate Towards Antibiotic Resistance.” JAC-Antimicrobial
Resistance 6(3):dlae085. https://doi.org/10.1093/jacamr/dlae085.
van Oldenborgh, G.J., E. Mitchell-Larson, G.A. Vecchi, H. de Vries, R. Vautard, and F. Otto. 2019.
“Cold Waves are Getting Milder in the Northern Midlatitudes.” Environmental Research Letters
14:114004. https://doi.org/10.1088/1748-9326/ab4867.
van Vliet, M.T.H., J. Thorslund, M. Strokal, N. Hofstra, M. Flörke, H. Ehalt Macedo, A. Nkwasa, et al.
2023. “Global River Water Quality Under Climate Change and Hydroclimatic Extremes.” Nature
Reviews Earth & Environment 4(10):687–702. https://doi.org/10.1038/s43017-023-00472-3.
Verhoeven, J.I., Y. Allach, I.C.H. Vaartjes, C.J.M. Klijn, and F.-E. De Leeuw. 2021. “Ambient Air
Pollution and the Risk of Ischaemic and Haemorrhagic Stroke.” The Lancet Planetary Health
5(8):e542–e552. https://doi.org/10.1016/S2542-5196(21)00145-5.
Vicedo-Cabrera, A.M., N. Scovronick, F. Sera, D. Royé, R. Schneider, A. Tobias, C. Astrom, et al. 2021.
“The Burden of Heat-Related Mortality Attributable to Recent Human-Induced Climate Change.”
Nature Climate Change 11(6):492–500. https://doi.org/10.1038/s41558-021-01058-x.
Vicente-Serrano, S.M., D. Peña-Angulo, S. Beguería, F. Domínguez-Castro, M. Tomás-Burguera, I.
Noguera, L. Gimeno-Sotelo, and A. El Kenawy. 2022. “Global Drought Trends and Future
Projections.” Philosophical Transactions of the Royal Society A: Mathematical, Physical and
Engineering Sciences 380(2238):20210285. https://doi.org/10.1098/rsta.2021.0285.
von Schuckmann, K., M.D. Palmer, K.E. Trenberth, A. Cazenave, D. Chambers, N. Champollion, J.
Hansen, et al. 2016. “An Imperative to Monitor Earth’s Energy Imbalance.” Nature Climate Change
6(2):138–144. https://doi.org/10.1038/nclimate2876.
Vongelis, P., N.G. Koulouris, P. Bakakos, and N. Rovina. 2025. “Air Pollution and Effects of
Tropospheric Ozone (O3) on Public Health.” International Journal of Environmental Research and
Public Health 22(5):709. https://doi.org/10.3390/ijerph22050709.
Vonk, J.E., S.E. Tank, W.B. Bowden, I. Laurion, W.F. Vincent, P. Alekseychik, M. Amyot, et al. 2015.
“Reviews and Syntheses: Effects of Permafrost Thaw on Arctic Aquatic Ecosystems.”
Biogeosciences 12:7129–7167. https://doi.org/10.5194/bg-12-7129-2015.
Waidelich, P., F. Batibeniz, J. Rising, J.S. Kikstra, and S.I. Seneviratne. 2024. “Climate Damage
Projections Beyond Annual Temperature.” Nature Climate Change 14(6):592–599.
https://doi.org/10.1038/s41558-024-01990-8.
Warziniack, T., M. Arabi, T.C. Brown, P. Froemke, R. Ghosh, S. Rasmussen, and R. Swartzentruber.
2022. “Projections of Freshwater Use in the United States Under Climate Change.” Earth’s Future
10(2):e2021EF002222. https://doi.org/10.1029/2021EF002222.
Wdowinski, S., R. Bray, B.P. Kirtman, and Z. Wu. 2016, “Increasing Flooding Hazard in Coastal
Communities due to Rising Sea Level: Case Study of Miami Beach, Florida.” Ocean & Coastal
Management 126:1–8. https://doi.org/10.1016/j.ocecoaman.2016.03.002.
Weeda, L.J.Z., C.J.A. Bradshaw, M.A. Judge, C.M. Saraswati, and P.N. Le Souëf. 2024. “How Climate
Change Degrades Child Health: A Systematic Review and Meta-Analysis.” The Science of the Total
Environment 920:170944. https://doi.org/10.1016/j.scitotenv.2024.170944.
Wei, J., J. Wang, Z. Li, S. Kondragunta, S. Anenberg, Y. Wang, H. Zhang, et al. 2023. “Long-Term
Mortality Burden Trends Attributed to Black Carbon and PM2·5 from Wildfire Emissions Across the
Prepublication Copy
104 Effects of Human-Caused GHG Emissions on U.S. Climate, Health, and Welfare
Continental USA from 2000 to 2020: A Deep Learning Modelling Study.” The Lancet Planetary
Health 7(12):e963–e975. https://doi.org/10.1016/S2542-5196(23)00235-8.
Weinberger, K.R., D. Harris, K.R. Spangler, A. Zanobetti, and G.A. Wellenius. 2020. “Estimating the
Number of Excess Deaths Attributable to Heat in 297 United States Counties.” Environmental
Epidemiology 4(3):e096. https://doi.org/10.1097/EE9.0000000000000096.
Weis, J.S., E.B. Watson, B. Ravit, C. Harman, and M. Yepsen. 2021. “The Status and Future of Tidal
Marshes in New Jersey Faced with Sea Level Rise.” Anthropocene Coasts 4(1):168–192.
https://doi.org/10.1139/anc-2020-0020.
West, J.J., C.G. Nolte, M.L. Bell, A.M. Fiore, P.G. Georgopoulos, J.J. Hess, L.J. Mickley, et al. 2023.
Chapter 14: Air Quality. Fifth National Climate Assessment. U.S. Global Change Research Program.
https://doi.org/10.7930/nca5.2023.ch14.
West, J.J., A.M. Fiore, L.W. Horowitz, and D.L. Mauzerall. 2006. “Global Health Benefits of Mitigating
Ozone Pollution with Methane Emission Controls.” Proceedings of the National Academy of Sciences
103(11):3988–3993. https://doi.org/10.1073/pnas.0600201103.
West, J.J., and A.M. Fiore. 2005. “Tropospheric Ozone by Reducing Methane Emissions.” Environmental
Science & Technology 39(13):4685–4691. https://doi.org/10.1021/es048629f.
Wettstein Z. S., C. Parrish, A. K. Sabbatini, M. H. Rogers, E. Seto, J. J. Hess. 2025. Emergency Care,
Hospitalization Rates, and Floods. JAMA Network Open 8(3):e250371.
https://doi.org/10.1001/jamanetworkopen.2025.0371.
Whitehead, J.C., E.L. Mecray, E.D. Lane, L. Kerr, M.L. Finucane, D.R. Reidmiller, M.C. Bove, et al.
2023. “Chapter 21: Northeast.” In Fifth National Climate Assessment. Crimmins, A.R., C.W. Avery,
D.R. Easterling, K.E. Kunkel, B.C. Stewart, and T.K. Maycock, eds. U.S. Global Change Research
Program, Washington, DC, USA. https://doi.org/10.7930/NCA5.2023.CH21.
Wijkström, J., R. Leiva, C.-G. Elinder, S. Leiva, Z. Trujillo, L. Trujillo, L., M. Söderberg, K. Hultenby,
and A. Wernerson. 2013. “Clinical and Pathological Characterization of Mesoamerican Nephropathy:
A New Kidney Disease in Central America.” American Journal of Kidney Diseases 62(5):908–918.
https://doi.org/10.1053/j.ajkd.2013.05.019.
Williams, E.L., C. Funk, and S. Shukla. 2023. “Anthropogenic Climate Change Negatively Impacts
Vegetation and Forage Conditions in the Greater Four Corner Region.” Earth’s Future
11(7):e2022EF002943. https://doi.org/10.1029/2022EF002943.
Williams, A.P., B.I. Cook, and J.E. Smerdon. 2022. “Rapid Intensification of the Emerging Southwestern
North American Megadrought.” Nature Climate Change 12:232–234. https://doi.org/10.1038/s41558-
022-01290-z.
Wolfe, D.W., R.M. Gifford, D. Hilbert, and Y. Luo. 1998. “Integration of Photosynthetic Acclimation to
CO2 at the Whole‐Plant Level.” Global Change Biology 4(8):879–893.
https://doi.org/10.1046/j.1365-2486.1998.00183.x.
Woolway, R.I., S. Sharma, and J.P. Smol. 2022. “Lakes in Hot Water: The Impacts of a Changing
Climate on Aquatic Ecosystems.” BioScience 72(11):1050–1061.
https://doi.org/10.1093/biosci/biac052.
World Health Organization (WHO). 2009. WHO Guidelines for Indoor Air Quality: Dampness and
Mould. WHO Regional Office for Europe. https://iris.who.int/handle/10665/164348
WHO. 2025. “Climate Change, Air Pollution, Pollen and Health: Technical Brief.” World Health
Organization. https://doi.org/10.2471/B09412.
WHO. 2018. COP24 Special Report: Health and Climate Change. https://www.who.int/publications/i/
item/9789241514972.
Wright, R.J., and J.G. Demain. 2024. “Growing Impact of Climate Change on Respiratory Health and
Related Allergic Disorders: Need for Health Systems to Prepare.” Immunology and Allergy Clinics of
North America 44(1):xi–xv. https://doi.org/10.1016/j.iac.2023.10.001.
Wu, Y., B. Wen, D. Gasevic, J.A. Patz, A. Haines, K.L. Ebi, V. Murray, S. Li, and Y. Guo. 2024.
“Climate Change, Floods, and Human Health.” New England Journal of Medicine 391(20):1949–
1958. https://doi.org/10.1056/NEJMsr2402457.
Prepublication Copy
References 105
Wunderling, N., A.S. von der Heydt, Y. Aksenov, S. Barker, R. Bastiaansen, V. Brovkin, M. Brunetti, et
al. 2024. “Climate Tipping Point Interactions and Cascades: A Review.” Earth System Dynamics
15(1):41–74. https://doi.org/10.5194/esd-15-41-2024.
Xu, J., N.M. Niehoff, A.J. White, E.J. Werder, and D.P. Sandler. 2022. “Fossil-Fuel and Combustion-
Related Air Pollution and Hypertension in the Sister Study.” Environmental Pollution 315:120401.
https://doi.org/10.1016/j.envpol.2022.120401.
Xu, R., P. Yu, M.J. Abramson, F.H. Johnston, J.M. Samet, M.L. Bell, A. Haines, K.L. Ebi, S. Li, and Y.
Guo. 2020. “Wildfires, Global Climate Change, and Human Health.” New England Journal of
Medicine 383(22):2173–2181. https://doi.org/10.1056/NEJMsr2028985.
Ye, X., R. Wolff, W. Yu, P. Vaneckova, X. Pan, and S. Tong. 2012. “Ambient Temperature and
Morbidity: A Review of Epidemiological Evidence.” Environmental Health Perspectives 120(1):19–
28. https://doi.org/10.1289/ehp.1003198.
Young, R., and S. Hsiang. 2024. “Mortality Caused by Tropical Cyclones in the United States.” Nature
635(8037):121–128. https://doi.org/10.1038/s41586-024-07945-5.
Yu, Y., O.V. Kalashnikova, M.J. Garay, and M. Notaro. 2019. “Climatology of Asian Dust Activation
and Transport Potential Based on MISR Satellite Observations and Trajectory Analysis.”
Atmospheric Chemistry and Physics 19:363–378. https://doi.org/10.5194/acp-19-363-2019.
Zanobetti, A., and A. Peters. 2014. “Disentangling Interactions between Atmospheric Pollution and
Weather.” Journal of Epidemiology and Community Health 69(7):613–615.
https://doi.org/10.1136/jech-2014-203939.
Zelinka, M.D., T.A. Myers, D.T. McCoy, S. Po-Chedley, P.M. Caldwell, P. Ceppi, S.A. Klein, and K.E.
Taylor. 2020. “Causes of Higher Climate Sensitivity in CMIP6 Models.” Geophysical Research
Letters 47(1)e2019GL085782.
https://doi.org/10.1029/1029GL085782https://doi.org/10.1029/2019GL085782.
Zemp, M., M. Huss, E. Thibert, N. Eckert, R. McNabb, J. Huber, M. Barandun, et al. 2019. “Global
Glacier Mass Changes and Their Contributions to Sea-Level Rise from 1961 to 2016.” Nature
568:382–386. https://doi.org/10.1038/s41586-019-1071-0.
Zeng, X., P. Broxton, and N. Dawson. 2018. “Snowpack Change From 1982 to 2016 Over Conterminous
United States.” Geophysical Research Letters 45(23). https://doi.org/10.1029/2018GL079621.
Zhang, L., T.L. Delworth, V. Koul, A. Ross, C. Stock, X. Yang, F. Zeng, et al. 2025. “Skillful Multiyear
Prediction of Flood Frequency Along the US Northeast Coast Using a High-Resolution Modeling
System.” Science Advances 11(20):eads4419. https://doi.org/10.1126/sciadv.ads4419.
Zhang, J., S. Bai, S. Lin, L. Cui, X. Zhao, S. Du, and Z. Wang. 2023. “Maternal Apparent Temperature
During Pregnancy on the Risk of Offspring Asthma and Wheezing: Effect, Critical Window, and
Modifiers.” Environmental Science and Pollution Research 30(22):62924–62937.
https://doi.org/10.1007/s11356-023-26234-8.
Zhang, S., L. Zhou, L. Zhang, Y. Yang, Z. Wei, S. Zhou, D. Yang, et al. 2022. “Reconciling
Disagreement on Global River Flood Changes in a Warming Climate.” Nature Climate Change
12(12):1160–1167. https://doi.org/10.1038/s41558-022-01539-7.
Zhang, Y., and A.L. Steiner. 2022. “Projected Climate-Driven Changes in Pollen Emission Season
Length and Magnitude over the Continental United States.” Nature Communications 13(1):1234.
https://doi.org/10.1038/s41467-022-28764-0.
Zhang, F., J.A. Biederman, M.P. Dannenberg, D. Yan, S.C. Reed, and W.K. Smith. 2021. “Five Decades
of Observed Daily Precipitation Reveal Longer and More Variable Drought Events Across Much of
the Western United States.” Geophysical Research Letters 48(7):e2020GL092293.
https://doi.org/10.1029/2020GL092293.
Zhang, Y., O.R. Cooper, A. Gaudel, P. Nédélec, S.-Y. Ogino, A.M. Thompson, and J.J. West. 2016.
“Tropospheric Ozone Change from 1980 to 2010 Dominated by Equatorward Redistribution of
Emissions.” Nature Geoscience 9(12):875–879. https://doi.org/10.1038/NGEO2827.
Prepublication Copy
106 Effects of Human-Caused GHG Emissions on U.S. Climate, Health, and Welfare
Zhao, H., L. Zhang, M.B. Kirkham, S.M. Welch, J.W. Nielsen-Gammon, G. Bai, J. Luo, et al. 2022.
“U.S. Winter Wheat Yield Loss Attributed to Compound Hot-Dry-Windy Events.” Nature
Communications 13(1):7233. https://doi.org/10.1038/s41467-022-34947-6.
Zhiyuan, M., Y. Liu, J. Liu, and Y. Lyu. 2025. “Mitigation of Thermal Effects in Bridges: A
Comprehensive Review of Control Methodologies.” Journal of Traffic and Transportation
Engineering 12(2):215–235. https://doi.org/10.1016/j.jtte.2024.12.003.
Zhou, X., K. Josey, L. Kamareddine, M.C. Caine, T. Liu, L.J. Mickley, M. Cooper, and F. Dominici.
2021. “Excess of COVID-19 Cases and Deaths Due to Fine Particulate Matter Exposure During the
2020 Wildfires in the United States.” Science Advances 7(33):eabi8789.
https://doi.org/10.1126/sciadv.abi8789.
Zhu, P., J. Burney, J. Chang, Z. Jin, N.D. Mueller, Q. Xin, J. Xu, J., L. Yu, D. Makowski, and P. Ciais.
2022. “Warming Reduces Global Agricultural Production by Decreasing Cropping Frequency and
Yields.” Nature Climate Change 12(11):1016–1023. https://doi.org/10.1038/s41558-022-01492-5.
Zhu, T., C.F. Fonseca De Lima, and I. De Smet. 2021. “Heat is On: How Crop Growth, Development,
and Yield Respond to High Temperature.” Journal of Experimental Botany 72(21):7359–7373.
https://doi.org/10.1093/jxb/erab308.
Zhuang, Y., R. Fu, J. Lisonbee, A.M. Sheffield, B.A. Parker, and G. Deheza. 2024. “Anthropogenic
Warming Has Ushered in an Era of Temperature-Dominated Droughts in the Western United States.”
Science Advances 10(45):eadn9389. https://doi.org/10.1126/sciadv.adn9389.
Ziska, L., K. Knowlton, C. Rogers, D. Dalan, N. Tierney, M.A. Elder, W. Filley, et al. 2011. “Recent
Warming by Latitude Associated with Increased Length of Ragweed Pollen Season in Central North
America.” Proceedings of the National Academy of Sciences 108(10):4248-4251.
https://doi.org/10.1073/pnas.1014107108.
Ziska, L.H. 2003. “Evaluation of the Growth Response of Six Invasive Species to Past, Present and
Future Atmospheric Carbon Dioxide.” Journal of Experimental Botany 54(381):395–404.
https://doi.org/10.1093/jxb/erg027
Ziska, L.H., J.R. Teasdale, and J.A. Bunce. 1999. “Future Atmospheric Carbon Dioxide May Increase
Tolerance to Glyphosate.” Weed Science 47(5):608–615. https://doi.org/10.1017/S0043174500092341.
Zou, C.-Z., and H. Qian. 2016. “Stratospheric Temperature Climate Data Record from Merged SSU and
AMSU-A Observations.” Journal of Atmospheric and Oceanic Technology 33(9):1967–1984.
https://doi.org/10.1175/jtech-d-16-0018.1.
Zscheischler, J., S. Westra, B.J.J.M. van den Hurk, S.I. Seneviratne, P.J. Ward, A. Pitman, A.
Aghakouchak, et al. 2018. “Future Climate Risk from Compound Events.” Nature Climate Change
8:469–477. https://doi.org/10.1038/s41558-018-0156-3.
Prepublication Copy
Appendix A
Committee Member Biographies
Shirley Tilghman (NAS/NAM) (Chair) is a Professor of Molecular Biology and Public Affairs Emerita
at Princeton University, where she served as the 19th president from 2001 to 2013. Following her
retirement from the presidency she returned to the faculty, where she had been teaching and conducting
research since 1986 as an Investigator of the Howard Hughes Medical Institute. She is best known for her
pioneering work on genomic imprinting. Tilghman is a recipient of the L’Oréal-UNESCO Award for
Women in Science, the Lifetime Achievement Award from the Society for Developmental Biology, the
Genetics Society of America Medal for outstanding contributions to her field, and the society’s George
W. Beadle Award for contribution to the genetics community. Tilghman is a member of the National
Academy of Sciences, the National Academy of Medicine, the American Philosophical Society, the Royal
Society of London, and is an Officer of the Order of Canada. She serves as a trustee of the Broad Institute
of the Massachusetts Institute of Technology and Harvard University, the Institute for Advanced Study,
the Simons Foundation and the Hypothesis Fund. Tilghman received an Honors B.Sc. in chemistry from
Queen’s University at Kingston, Canada, and a Ph.D. in biochemistry from Temple University. She
previously served on the National Academies’ Committee on Mapping and Sequencing the Human
Genome, the Board on Life Sciences, and the Roundtable on Aligning Incentives for Open Science.
David Allen (NAE) is the Norbert Dittrich-Welch Chair in Chemical Engineering and the co-Director of
the Center for Energy and Environmental Systems Analyses at the University of Texas at Austin. His
expertise is in urban air quality and the engineering of sustainable systems. He has been a lead
investigator for multiple air quality measurement and modeling studies, including studies that reported
some of the first measurements of methane emissions from oil and gas supply chains. He is a member of
the National Academy of Engineering and a recipient of the ENI Energy Transition Award. He previously
served on the U.S. Environmental Protection Agency Science Advisory Board (EPA SAB) and the U.S.
Department of Energy National Petroleum Council. Allen received a B.S. degree in chemical engineering,
with distinction, from Cornell University, and an M.S. and Ph.D. in chemical engineering from the
California Institute of Technology. He previously served on the National Academies’ Committee on
Anthropogenic Methane Emissions in the United States, the Board on Energy and Environmental
Systems, and chaired the Committee on the Chemistry of Urban Wildfires.
Allen has previously provided consulting services to oil and gas companies and consortia
involving oil and gas companies. The Center for Energy and Environmental Systems Analyses is
supported by research gifts through an Industrial Affiliates Program. During his service on the EPA SAB
he co-authored public reports and made public statements related to methane emission regulations under
consideration by the EPA.
Susan Anenberg is a Professor and Chair of the Environmental and Occupational Health Department at
the George Washington (GW) University Milken Institute School of Public Health. She is also the
Director of the GW Climate and Health Institute. Previously, she was a Co-Founder and Partner at
Environmental Health Analytics, LLC, the Deputy Managing Director for Recommendations at the U.S.
Chemical Safety Board, an environmental scientist at the U.S. Environmental Protection Agency (EPA),
and a senior advisor for clean cookstove initiatives at the U.S. State Department. Her research focuses on
the health implications of air pollution and climate change, from local to global scales. She currently
serves on the World Health Organization’s Global Air Pollution and Health Technical Advisory Group
and previously served on the EPA Science Advisory Board and as President of the GeoHealth section of
the American Geophysical Union. Anenberg received a B.A. in biology and environmental science from
108 Effects of Human-Caused GHG Emissions on U.S. Climate, Health, and Welfare
Northwestern University and a Ph.D. in environmental science and engineering and environmental policy
from the University of North Carolina. She previously served on the National Academies’ Committee on
Utilizing Advanced Environmental Health and Geospatial Data and Technologies to Inform Community
Investment and the Committee to Advise the U.S. Global Change Research Program.
Anenberg has previously submitted public comments on proposed EPA rules related to
greenhouse gas emissions and air pollution. She receives research support for work on climate and health
from the Natural Resources Defense Council. Annenberg signed a brief of amici curiae submitted to the
Supreme Court in State of West Virginia, et al., v. U.S. Environmental Protection Agency, et al., that
argued anthropogenic climate change, fueled by emissions of greenhouse gases such as carbon dioxide,
harms public health in the United States.
Michele Barry (NAM) is the Drs. Ben & A. Jess Shenson Professor of Medicine and Tropical Diseases
at Stanford University. She is also the Director of the Center for Innovation in Global Health and Senior
Associate Dean for Global Health. Prior to her current role, she was a Professor of Medicine at Yale from
which she was recruited to be the inaugural dean for global health at Stanford. She has published in the
areas of climate’s impact on health, tropical diseases, and human and planetary health. She is a member of
the National Academy of Medicine (NAM), the Council on Foreign Relations, and the American
Academy of Arts and Sciences. Barry is Chair Emerita of the Board of Directors for the Consortium of
Universities for Global Health and a past President of the American Society of Tropical Medicine and
Hygiene (ASTMH). She is a recipient of the Ben Kean Medal from the ASTMH and the Elizabeth
Blackburn Award from the American Medical Woman’s Association. Barry received an A.B. from Bryn
Mawr College and an M.D. from the Albert Einstein College of Medicine. She previously served on the
National Academies’ Board on Global Health and has co-led the NAM climate interest group.
Susan Hanson (NAS) is an urban geographer, now retired for nearly 20 years from Clark University,
where she taught for 25 years and was the Landry Professor of Geography and Director of the School of
Geography. Prior to Clark, she was an Assistant and Associate Professor at SUNY Buffalo. Her research
has focused on the relationship between people and the urban built environment. This has included
understanding people’s everyday travel-activity patterns, examining the way that different groups make
use of the city and showing how urban spatial structure configures household travel. She also seeks to
understand the emergence of sustainable versus unsustainable practices in urban areas. Hanson’s Awards
include the Lifetime Achievement Award and the Award for Creativity, both from the American
Prepublication Copy
Appendix A 109
Association of Geographers, as well as the Carey Award for Leadership & Service from the
Transportation Research Board of the National Academies. She is a member of the American Academy of
Arts and Sciences and the National Academy of Sciences. Hanson received an A.B. in geography from
Middlebury College and an M.S. and Ph.D. in geography from Northwestern University. Prior to graduate
school, she was a Peace Corps Volunteer in Western Kenya, teaching at a boys’ secondary school.
Chris Hendrickson (NAE) is the Hamerschlag University Professor of Engineering Emeritus and
director of the Traffic 21 Institute at Carnegie Mellon University. His expertise is in engineering planning
and management, including transportation systems, design for the environment, system performance,
construction project management, finance, and computer applications. Central themes in his work are a
systems-wide perspective and a balance of engineering and management considerations. He is the editor-
in-chief of the American Society of Civil Engineers (ASCE) Journal of Transportation Engineering Part A
(Systems). Hendrickson is a recipient of the Council of University Transportation Centers Lifetime
Achievement Award, the American Road & Transportation Builders Association Steinburg Award, the
Faculty Award of the Carnegie Mellon Alumni Association, the Turner Lecture Award of the ASCE, and
the Fenves Systems Research Award from the Institute of Complex Engineering Systems. He is a member
of the National Academy of Engineering, the National Academy of Construction, a fellow of the
American Association for the Advancement of Science, and a distinguished member of the ASCE.
Hendrickson received a B.S. in general engineering (resources strategy) and an M.S. in civil engineering
from Stanford University, an M.Phil. in economics from Oxford University, and a Ph.D. in civil
engineering from the Massachusetts Institute of Technology. He previously served on the National
Academies’ Committee on Accelerating Decarbonization in the United States: Technology, Policy, and
Societal Dimensions and chaired the Transportation Research Board Division Committee.
Marika Holland is a Senior Scientist at the National Science Foundation National Center for
Atmospheric Research (NSF NCAR). Her research is focused on polar climate variability and change.
She has extensive experience in using climate models to study coupled climate interactions and has been
active in the development of improved sea ice models for climate simulations. Holland has served as co-
chair for the Polar Climate Working Group of the Community Earth System Model and Chief Scientist
for the Community Earth System Model project. She contributed to the Third, Fourth, Fifth, and Sixth
Assessment Reports of the Intergovernmental Panel on Climate Change. Holland is a fellow of the
American Geophysical Union and the American Meteorological Society and a recipient of the
International Arctic Science Committee Medal and the Community Earth System Model Distinguished
Achievement Award. Holland received a Ph.D. in atmosphere and ocean sciences from the University of
Colorado and was a Postdoctoral Fellow at the University of Victoria in British Columbia. She previously
served on the National Academies’ Committee on Understanding and Monitoring Abrupt Climate
Change.
Prepublication Copy
110 Effects of Human-Caused GHG Emissions on U.S. Climate, Health, and Welfare
Arthur Lee retired from Chevron in 2024, where he was a Chevron Fellow. During his Chevron career
he held roles of increasing responsibilities, including the corporation-wide formulation of strategic
positioning and policy development on issues ranging from Chevron’s internal energy policy to U.S. air
pollution issues and actions addressing climate change concerns. He continues to mentor employees as
Chevron Fellow Emeritus in retirement. Lee represented Chevron in numerous roles: Chair of the
International Petroleum Industry Environmental Conservation Association’s Climate Change Working
Group, member of the board of directors of the International Emissions Trading Association, and member
of the Executive Committee of the International Energy Agency Greenhouse Gas R&D Programme. Prior
to Chevron, Lee held positions as an engineer with the U.S. Environmental Protection Agency, Fluor
Daniel Inc., and Directed Technologies. He served on the Intergovernmental Panel on Climate Change
(IPCC) Fourth Assessment as Review Editor in the Special Report on Carbon Dioxide Capture and
Storage, contributing author on geothermal energy in the Special Report on Renewable Energy, and again
as Review Editor in the Sixth Assessment, and he was awarded a certificate recognizing his contribution
to the IPCC’s Nobel Peace Prize. Additionally, Lee served on the National Climate Assessment
Development and Advisory Committee and was a coordinating lead author of the climate change
adaptation chapter. Lee received a B.S. in chemical engineering from the Massachusetts Institute of
Technology and an M.S. in chemical engineering from the California Institute of Technology. He
previously served on the National Academies’ Board on Atmospheric Sciences and Climate.
Kari Nadeau (NAM) is the Chair of the Department of Environmental Health and the John Rock
Professor of Climate Science and Population Studies at the Harvard T.H. Chan School of Public Health.
She is also a Professor of Medicine at Harvard Medical School and works at the Beth Israel Deaconess
Medical Center. Her expertise is in immunology, allergies and asthma, and climate change solutions, with
a focus on understanding how environmental and epigenetic factors affect the risk of developing immune
dysfunction. Her wet lab laboratory has been studying exposomics and solutions-facing research with
policy-oriented outcomes. Nadeau has started four biotechnology companies, co-started a sustainability
seed grant program, and works with the World Health Organization and United Nations on several
projects in environmental and global health. She is a member of the National Academy of Medicine, the
American Association of Physicians, the American Society of Clinical Investigation, and a Fellow of the
American Academy of Allergy, Asthma, and Immunology. Nadeau received a degree in biology from
Haverford College and an M.D. and Ph.D. in biological chemistry and molecular pharmacology from
Harvard Medical School. She completed a residency in pediatrics and a fellowship in allergy, asthma, and
immunology.
Nadeau has previously submitted a public comment on proposed EPA rules related to greenhouse
gas emissions and air pollution. She signed a brief of amici curiae submitted to the U.S. Court of Appeals
Central District of California in G.B., et al., v. U.S. Environmental Protection Agency, et al., that argued
that children are uniquely vulnerable to the consequences of rising temperatures and to increased air
pollution from climate change.
Charles Rice is a University Distinguished Professor in soil microbiology and holds the Vanier
University Professorship in the Department of Agronomy at Kansas State University. His research
focuses on soil carbon and nitrogen, soil health, microbial ecology, and the impacts of climate change on
agricultural and grassland ecosystems. Rice contributed to the Third and Fourth Assessment Reports of
the Intergovernmental Panel on Climate Change (IPCC) and he was awarded a certificate recognizing his
contribution to the IPCC’s Nobel Peace Prize. He is a Fellow of the Soil Science Society of America, the
American Society of Agronomy, Sigma Xi, and the American Association for the Advancement of
Science and a National Associate of the National Academies. Rice received a B.S. in natural
environmental systems from Northern Illinois University and an M.S. in soil science and Ph.D. in soil
microbiology from the University of Kentucky. He previously served on the National Academies’ Board
on Agriculture and Natural Resources.
Prepublication Copy
Appendix A 111
Drew Shindell (NAS) is Nicholas Professor of Earth Science at Duke University. He was previously a
senior scientist at the National Aeronautics and Space Administration’s Goddard Institute for Space
Studies. His expertise is in modeling the impact of emissions changes, and his work has investigated how
the atmospheric chemical system has important effects on humans through pollutants such as smog or
particulates, through acid rain, and through stratospheric ozone change, and how climate can be altered by
greenhouse gases, solar variability, volcanic eruptions, aerosols, and ozone, and what impacts changes in
climate and air quality may have on society. Shindell contributed to the Fifth Assessment Report of the
Intergovernmental Panel on Climate Change and the Fifth National Climate Assessment. He is a member
of the National Academy of Sciences and is a Fellow of the American Geophysical Union and the
American Association for the Advancement of Science. Shindell received a B.A. in physics from the
University of California, Berkeley and a Ph.D. in physics from the State University of New York Stony
Brook. He previously served on the National Academies’ Committee on the Provisions in the Internal
Revenue Code on Greenhouse Gas Emissions and the Committee on Assessment of Himalayan Glaciers:
Climate Change, Water Resources, and Water Security.
Shindell has previously submitted a public comment on proposed EPA rules related to
greenhouse gas emissions and air pollution. He has also provided Congressional testimony on the
relationship between climate and health as well as greenhouse gas emissions. Shindell signed briefs of
amici curiae submitted to the U.S. Court of Appeals District of Columbia Circuit in Competitive
Enterprise Institute, et al., v. National Highway Traffic and Safety Administration, et al., that argued that
the Safer Affordable Fuel Efficient Vehicles Rule would not adequately address fossil fuel emissions
from vehicles as well as a brief of amici curiae submitted to the U.S. Court of Appeals District of
Columbia Circuit in American Lung Association, et al., v. U.S. Environmental Protection Agency, et al.,
that argued that the Affordable Clean Energy Rule did not adequately consider impacts to historic
resources and communities.
Graeme Stephens (NAE) is the director of the Center for Climate Sciences at the Jet Propulsion
Laboratory (JPL). He was previously Distinguished University Professor at Colorado State University.
Stephens is the principal investigator of NASA’s CloudSat mission, and previously chaired the World
Climate Research Program Global Energy and Water and EXchanges (GEWEX) project, which examines
the topic of global water cycles and the global energy balance and the connections between. His research
activities focus on atmospheric radiation including the application of remote sensing in climate research
to understand the role of hydrological processes in climate change. Stephens is a member of the National
Academy of Engineering and a Fellow of the Royal Society, American Meteorological Society, American
Geophysical Union (AGU), and American Association for the Advancement of Science. He is a recipient
of the American Meteorological Society’s Houghton and Jule Charney Awards and the AGU Jule
Charney Lecturer. Stephens received a B.S. in physics and meteorology and a Ph.D. in meteorology from
the University of Melbourne, Australia. He previously served on the National Academies’ Committee on
Earth Science and Applications from Space.
David Titley is President and Founder of RV Weather, providing weather and routing services to the
Recreational Vehicle community. Previously, he was Professor of Practice in Meteorology and Professor
of International Affairs at The Pennsylvania State University. Titley served as a Naval Officer for 32
years and rose to the rank of Rear Admiral. Titley’s career included duties as Commander of the Naval
Meteorology and Oceanography Command, as well as Oceanographer and Navigator of the Navy. While
serving at the Pentagon, Titley initiated and led the U.S. Navy’s Task Force on Climate Change. After
retiring from the Navy, Titley served as the Deputy Undersecretary of Commerce for Operations, the
Chief Operating Officer position at the National Oceanic and Atmospheric Administration. Titley is an
expert on climate, the arctic, and national security. He received an honorary doctorate degree from the
University of Alaska Fairbanks and is a Fellow of the American Meteorological Society. Titley received a
B.S. in meteorology from Pennsylvania State University, an M.S. in meteorology and physical
Prepublication Copy
112 Effects of Human-Caused GHG Emissions on U.S. Climate, Health, and Welfare
oceanography and Ph.D. in meteorology from the Naval Postgraduate School. He previously served on
the National Academies’ Board on Atmospheric Sciences and Climate.
John Wall (NAE) retired from Cummins, Inc. in 2015, where he was Chief Technology Officer. He has
over 45 years of industry experience in the development of low-emission internal combustion engines and
fuels and working with regulatory agencies in the United States and worldwide to align engine and fuel
technologies with future emissions policy with the objective of delivering products meeting both
commercial and environmental expectations. During his time with Cummins, Wall was directly involved
in the most critical technology programs for low emissions, powertrain efficiency and alternative fuels
and he also led the growth of the Cummins technical organization from 1,000 engineers, mostly centered
in the United States to more than 6,000 engineers globally. Prior to joining Cummins, he led Diesel Fuels
Research for Chevron, where his team was first to discover the important contribution of fuel sulfur to
diesel particulate emissions, leading to the first low-sulfur fuel standards by the U.S. Environmental
Protection Agency (EPA) in 1994. Wall is a member of the National Academy of Engineering and a
Society of Automotive Engineers (SAE) Fellow. He received the SAE Horning and Colwell Awards for
research in the area of diesel fuel effects on emissions, SAE Pischinger Powertrain Innovation Award,
and ASME Honda Medal for significant contributions in the field of personal transportation, the
California Air Resources Board Haagen-Smit Award and EPA Zosel Award for career accomplishments
in diesel emission control, and has been recognized by the Health Effects Institute for technologic
innovation and commitment to clean air. Wall received an S.B., S.M., and Sc.D. in mechanical
engineering from the Massachusetts Institute of Technology. He currently serves on the National
Academies’ Board on Science, Technology, and Economic Policy and previously served on the Board on
Energy and Environmental Systems.
Prepublication Copy
Appendix B
Disclosure of Unavoidable Conflicts of Interest
The conflict of interest policy of the National Academies of Sciences, Engineering, and Medicine
(http://www.nationalacademies.org/coi) prohibits the appointment of an individual to a committee
authoring a Consensus Study Report if the individual has a conflict of interest that is relevant to the task
to be performed. An exception to this prohibition is permitted if the National Academies determine that
the conflict is unavoidable and the conflict is publicly disclosed. A determination of a conflict of interest
for an individual is not an assessment of that individual’s actual behavior or character or ability to act
objectively despite the conflicting interest.
Arthur Lee has a conflict of interest in relation to his service on the Committee on Anthropogenic
Greenhouse Gases and the U.S. Climate: Evidence and Impacts because he holds stock in Chevron
Corporation and has received reimbursement for travel expenses from Chevron incurred while mentoring
employees in his capacity as Chevron Fellow Emeritus.
The National Academies have concluded that for this committee to accomplish the tasks for
which it was established, its membership must include someone who has recent experience within the oil
and gas sector, particularly in developing corporate internal energy policy on air pollution issues. As
described in his biographical summary, Lee has extensive experience working at the intersection of
climate change and energy production, with expertise in oil and gas infrastructure and evaluating the
contributions of the oil and gas industry to climate change and strategies for reducing greenhouse gas
emissions.
The National Academies have determined that the experience and expertise of Lee are needed for
the committee to accomplish the task for which it has been established. The National Academies could
not find another available individual with the equivalent experience and expertise who does not have a
conflict of interest. Therefore, the National Academies have concluded that the conflict is unavoidable.
The National Academies believe that Lee can serve effectively as a member of the committee,
and the committee can produce an objective report, taking into account the composition of the committee,
the work to be performed, and the procedures to be followed in completing the study.
Wall has a conflict of interest in relation to his service on the Committee on Anthropogenic
Greenhouse Gases and the U.S. Climate: Evidence and Impacts because he owns stock in Cummins,
Shell, and Chevron.
The National Academies have concluded that for this committee to accomplish the tasks for
which it was established, its membership must include someone who has experience with the automotive
industry focusing on developing emissions policy with the objective of delivering products meeting both
commercial and environmental expectations. As described in his biographical summary, Wall has
extensive experience working at the intersection of climate change and the automotive industry, with
expertise in low-emission internal combustion engines and fuels and the environmental impacts of these
emissions sources.
The National Academies have determined that the experience and expertise of Wall are needed
for the committee to accomplish the task for which it has been established. The National Academies could
114 Effects of Human-Caused GHG Emissions on U.S. Climate, Health, and Welfare
not find another available individual with the equivalent experience and expertise who does not have a
conflict of interest. Therefore, the National Academies have concluded that the conflict is unavoidable.
The National Academies believe that Wall can serve effectively as a member of the committee,
and the committee can produce an objective report, taking into account the composition of the committee,
the work to be performed, and the procedures to be followed in completing the study.
Prepublication Copy
Appendix C
Executive Summary of the Technical Support Document
for Endangerment and Cause or Contribute Findings for
Greenhouse Gases under Section 202(a)
of the Clean Air Act (EPA, 2009)
EXECUTIVE SUMMARY
This document provides technical support for the endangerment and cause or contribute analyses
concerning greenhouse gas (GHG) emissions under section 202(a) of the Clean Air Act. This document
itself does not convey any judgment or conclusion regarding the question of whether GHGs may be
reasonably anticipated to endanger public health or welfare, as this decision is ultimately left to the
judgment of the Administrator. The conclusions here and the information throughout this document are
primarily drawn from the assessment reports of the Intergovernmental Panel on Climate Change (IPCC),
the U.S. Climate Change Science Program (CCSP), the U.S. Global Change Research Program
(USGCRP), and the National Research Council (NRC).
Greenhouse gases, once emitted, can remain in the atmosphere for decades to centuries, meaning
that 1) their concentrations become well-mixed throughout the global atmosphere regardless of
emission origin, and 2) their effects on climate are long lasting. The primary long-lived GHGs directly
emitted by human activities include carbon dioxide (CO2), methane (CH4), nitrous oxide (N2O),
hydrofluorocarbons (HFCs), perfluorocarbons (PFCs), and sulfur hexafluoride (SF6). Greenhouse gases
have a warming effect by trapping heat in the atmosphere that would otherwise escape to space.
In 2007, U.S. GHG emissions were 7,150 teragrams 1 of CO2 equivalent 2 (TgCO2eq). The dominant
gas emitted is CO2, mostly from fossil fuel combustion. Methane is the second largest component of
U.S. emissions, followed by N2O and the fluorinated gases (HFCs, PFCs, and SF6). Electricity generation
is the largest emitting sector (34% of total U.S. GHG emissions), followed by transportation (28%) and
industry (19%).
Transportation sources under Section 202 of the Clean Air Act (passenger cars, light duty trucks,
other trucks and buses, motorcycles, and cooling) emitted 1,649 TgCO2eq in 2007, representing
23% of total U.S. GHG emissions.
U.S. transportation sources under Section 202 made up 4.3% of total global GHG emissions in
2005, which, in addition to the United States as a whole, ranked only behind total GHG emissions from
1
One teragram (Tg) = 1 million metric tons. 1 metric ton = 1,000 kilograms = 1.102 short tons = 2,205 pounds.
2
Long-lived GHGs are compared and summed together on a CO2-equivalent basis by multiplying each gas by its
global warming potential (GWP), as estimated by IPCC. In accordance with United Nations Framework Convention
on Climate Change (UNFCCC) reporting procedures, the U.S. quantifies GHG emissions using the 100-year
timeframe values for GWPs established in the IPCC Second Assessment Report.
116 Effects of Human-Caused Greenhouse Gas Emissions on U.S. Climate, Health, & Welfare
China, Russia, and India but ahead of Japan, Brazil, Germany, and the rest of the world’s countries. In
2005, total U.S. GHG emissions were responsible for 18% of global emissions, ranking only behind
China, which was responsible for 19% of global GHG emissions.
U.S. emissions of sulfur oxides (SOx), nitrogen oxides (NOx), direct particulates, and ozone
precursors have decreased in recent decades, due to regulatory actions and improvements in
technology. Sulfur dioxide (SO2) emissions in 2007 were 5.9 Tg of sulfur, primary fine particulate matter
(PM2.5) emissions in 2005 were 5.0 Tg, NOx emissions in 2005 were 18.5 Tg, volatile organic compound
(VOC) emissions in 2005 were 16.8 Tg, and ammonia emissions in 2005 were 3.7 Tg.
The global atmospheric CO2 concentration has increased about 38% from pre-industrial levels to
2009, and almost all of the increase is due to anthropogenic emissions. The global atmospheric
concentration of CH4 has increased by 149% since pre-industrial levels (through 2007); and the N2O
concentration has increased by 23% (through 2007). The observed concentration increase in these gases
can also be attributed primarily to anthropogenic emissions. The industrial fluorinated gases, HFCs,
PFCs, and SF6, have relatively low atmospheric concentrations but the total radiative forcing due to these
gases is increasing rapidly; these gases are almost entirely anthropogenic in origin.
Historic data show that current atmospheric concentrations of the two most important directly
emitted, long-lived GHGs (CO2 and CH4) are well above the natural range of atmospheric
concentrations compared to at least the last 650,000 years. Atmospheric GHG concentrations have
been increasing because anthropogenic emissions have been outpacing the rate at which GHGs are
removed from the atmosphere by natural processes over timescales of decades to centuries.
Current ambient air concentrations of CO2 and other GHGs remain well below published exposure
thresholds for any direct adverse health effects, such as respiratory or toxic effects.
The global average net effect of the increase in atmospheric GHG concentrations, plus other human
activities (e.g., land-use change and aerosol emissions), on the global energy balance since 1750 has
been one of warming. This total net heating effect, referred to as forcing, is estimated to be +1.6 (+0.6 to
+2.4) watts per square meter (W/m2), with much of the range surrounding this estimate due to
uncertainties about the cooling and warming effects of aerosols. However, as aerosol forcing has more
regional variability than the well-mixed, long-lived GHGs, the global average might not capture some
regional effects. The combined radiative forcing due to the cumulative (i.e., 1750 to 2005) increase in
atmospheric concentrations of CO2, CH4, and N2O is estimated to be +2.30 (+2.07 to +2.53) W/m2. The
rate of increase in positive radiative forcing due to these three GHGs during the industrial era is very
likely to have been unprecedented in more than 10,000 years.
Warming of the climate system is unequivocal, as is now evident from observations of increases in
global average air and ocean temperatures, widespread melting of snow and ice, and rising global
average sea level. Global mean surface temperatures have risen by 1.3 ± 0.32°F (0.74°C ± 0.18°C) over
the last 100 years. Eight of the 10 warmest years on record have occurred since 2001. Global mean
surface temperature was higher during the last few decades of the 20th century than during any
comparable period during the preceding four centuries.
Most of the observed increase in global average temperatures since the mid-20th century is very
likely due to the observed increase in anthropogenic GHG concentrations. Climate model simulations
suggest natural forcing alone (i.e., changes in solar irradiance) cannot explain the observed warming.
Prepublication Copy
Appendix C 117
U.S. temperatures also warmed during the 20th and into the 21st century; temperatures are now
approximately 1.3°F (0.7°C) warmer than at the start of the 20th century, with an increased rate of
warming over the past 30 years. Both the IPCC and the CCSP reports attributed recent North American
warming to elevated GHG concentrations. In the CCSP (2008g) report, the authors find that for North
America, “more than half of this warming [for the period 1951-2006] is likely the result of human-caused
greenhouse gas forcing of climate change.”
Observations show that changes are occurring in the amount, intensity, frequency and type of
precipitation. Over the contiguous United States, total annual precipitation increased by 6.1% from 1901
to 2008. It is likely that there have been increases in the number of heavy precipitation events within
many land regions, even in those where there has been a reduction in total precipitation amount,
consistent with a warming climate.
There is strong evidence that global sea level gradually rose in the 20th century and is currently
rising at an increased rate. It is not clear whether the increasing rate of sea level rise is a reflection of
short-term variability or an increase in the longer-term trend. Nearly all of the Atlantic Ocean shows sea
level rise during the last 50 years with the rate of rise reaching a maximum (over 2 millimeters [mm] per
year) in a band along the U.S. east coast running east-northeast.
Satellite data since 1979 show that annual average Arctic sea ice extent has shrunk by 4.1% per
decade. The size and speed of recent Arctic summer sea ice loss is highly anomalous relative to the
previous few thousands of years.
Widespread changes in extreme temperatures have been observed in the last 50 years across all
world regions, including the United States. Cold days, cold nights, and frost have become less frequent,
while hot days, hot nights, and heat waves have become more frequent.
Observational evidence from all continents and most oceans shows that many natural systems are
being affected by regional climate changes, particularly temperature increases. However, directly
attributing specific regional changes in climate to emissions of GHGs from human activities is difficult,
especially for precipitation.
Ocean CO2 uptake has lowered the average ocean pH (increased acidity) level by approximately 0.1
since 1750. Consequences for marine ecosystems can include reduced calcification by shell-forming
organisms, and in the longer term, the dissolution of carbonate sediments.
Observations show that climate change is currently affecting U.S. physical and biological systems in
significant ways. The consistency of these observed changes in physical and biological systems and the
observed significant warming likely cannot be explained entirely due to natural variability or other
confounding non-climate factors.
Projections of Future Climate Change With Continued Increases in Elevated GHG Concentrations
Most future scenarios that assume no explicit GHG mitigation actions (beyond those already
enacted) project increasing global GHG emissions over the century, with climbing GHG
concentrations. Carbon dioxide is expected to remain the dominant anthropogenic GHG over the course
of the 21st century. The radiative forcing associated with the non-CO2 GHGs is still significant and
increasing over time.
Prepublication Copy
118 Effects of Human-Caused Greenhouse Gas Emissions on U.S. Climate, Health, & Welfare
Future warming over the course of the 21st century, even under scenarios of low-emission growth,
is very likely to be greater than observed warming over the past century. According to climate model
simulations summarized by the IPCC, through about 2030, the global warming rate is affected little by the
choice of different future emissions scenarios. By the end of the 21st century, projected average global
warming (compared to average temperature around 1990) varies significantly depending on the emission
scenario and climate sensitivity assumptions, ranging from 3.2 to 7.2ºF (1.8 to 4.0ºC), with an uncertainty
range of 2.0 to 11.5ºF (1.1 to 6.4ºC).
All of the United States is very likely to warm during this century, and most areas of the United
States are expected to warm by more than the global average. The largest warming is projected to
occur in winter over northern parts of Alaska. In western, central and eastern regions of North America
the projected warming has less seasonal variation and is not as large, especially near the coast, consistent
with less warming over the oceans.
It is very likely that heat waves will become more intense, more frequent, and longer lasting in a
future warm climate, whereas cold episodes are projected to decrease significantly.
Increases in the amount of precipitation are very likely in higher latitudes, while decreases are
likely in most subtropical latitudes and the southwestern United States, continuing observed
patterns. The mid-continental area is expected to experience drying during summer, indicating a greater
risk of drought.
Intensity of precipitation events is projected to increase in the United States and other regions of
the world. More intense precipitation is expected to increase the risk of flooding and result in greater
runoff and erosion that has the potential for adverse water quality effects.
It is likely that hurricanes will become more intense, with stronger peak winds and more heavy
precipitation associated with ongoing increases of tropical sea surface temperatures. Frequency changes
in hurricanes are currently too uncertain for confident projections.
By the end of the century, global average sea level is projected by IPCC to rise between 7.1 and 23
inches (18 and 59 centimeter [cm]), relative to around 1990, in the absence of increased dynamic ice
sheet loss. Recent rapid changes at the edges of the Greenland and West Antarctic ice sheets show
acceleration of flow and thinning. While an understanding of these ice sheet processes is incomplete, their
inclusion in models would likely lead to increased sea level projections for the end of the 21st century.
Sea ice extent is projected to shrink in the Arctic under all IPCC emissions scenarios.
Risk to society, ecosystems, and many natural Earth processes increase with increases in both the
rate and magnitude of climate change. Climate warming may increase the possibility of large,
abrupt regional or global climatic events (e.g., disintegration of the Greenland Ice Sheet or collapse
of the West Antarctic Ice Sheet). The partial deglaciation of Greenland (and possibly West Antarctica)
could be triggered by a sustained temperature increase of 2 to 7ºF (1 to 4ºC) above 1990 levels. Such
warming would cause a 13 to 20 feet (4 to 6 meter) rise in sea level, which would occur over a time
period of centuries to millennia.
Prepublication Copy
Appendix C 119
CCSP reports that climate change has the potential to accentuate the disparities already evident in
the American health care system, as many of the expected health effects are likely to fall
disproportionately on the poor, the elderly, the disabled, and the uninsured. IPCC states with very
high confidence that climate change impacts on human health in U.S. cities will be compounded by
population growth and an aging population.
Severe heat waves are projected to intensify in magnitude and duration over the portions of the
United States where these events already occur, with potential increases in mortality and morbidity,
especially among the elderly, young, and frail.
Some reduction in the risk of death related to extreme cold is expected. It is not clear whether
reduced mortality from cold will be greater or less than increased heat-related mortality in the United
States due to climate change.
Increases in regional ozone pollution relative to ozone levels without climate change are expected
due to higher temperatures and weaker circulation in the United States and other world cities
relative to air quality levels without climate change. Climate change is expected to increase regional
ozone pollution, with associated risks in respiratory illnesses and premature death. In addition to human
health effects, tropospheric ozone has significant adverse effects on crop yields, pasture and forest
growth, and species composition. The directional effect of climate change on ambient particulate matter
levels remains uncertain.
Within settlements experiencing climate change, certain parts of the population may be especially
vulnerable; these include the poor, the elderly, those already in poor health, the disabled, those living
alone, and/or indigenous populations dependent on one or a few resources. Thus, the potential impacts of
climate change raise environmental justice issues.
CCSP concludes that, with increased CO2 and temperature, the life cycle of grain and oilseed crops
will likely progress more rapidly. But, as temperature rises, these crops will increasingly begin to
experience failure, especially if climate variability increases and precipitation lessens or becomes
more variable. Furthermore, the marketable yield of many horticultural crops (e.g., tomatoes, onions,
fruits) is very likely to be more sensitive to climate change than grain and oilseed crops.
Higher temperatures will very likely reduce livestock production during the summer season in
some areas, but these losses will very likely be partially offset by warmer temperatures during the
winter season.
Cold-water fisheries will likely be negatively affected; warm-water fisheries will generally benefit;
and the results for cool-water fisheries will be mixed, with gains in the northern and losses in the
southern portions of ranges.
Climate change has very likely increased the size and number of forest fires, insect outbreaks, and
tree mortality in the interior West, the Southwest, and Alaska, and will continue to do so. Over
North America, forest growth and productivity have been observed to increase since the middle of the
20th century, in part due to observed climate change. Rising CO2 will very likely increase photosynthesis
for forests, but the increased photosynthesis will likely only increase wood production in young forests on
fertile soils. The combined effects of expected increased temperature, CO2, nitrogen deposition, ozone,
and forest disturbance on soil processes and soil carbon storage remain unclear.
Coastal communities and habitats will be increasingly stressed by climate change impacts
interacting with development and pollution. Sea level is rising along much of the U.S. coast, and the
Prepublication Copy
120 Effects of Human-Caused Greenhouse Gas Emissions on U.S. Climate, Health, & Welfare
rate of change will very likely increase in the future, exacerbating the impacts of progressive inundation,
storm-surge flooding, and shoreline erosion. Storm impacts are likely to be more severe, especially along
the Gulf and Atlantic coasts. Salt marshes, other coastal habitats, and dependent species are threatened by
sea level rise, fixed structures blocking landward migration, and changes in vegetation. Population growth
and rising value of infrastructure in coastal areas increases vulnerability to climate variability and future
climate change.
Climate change will likely further constrain already overallocated water resources in some regions
of the United States, increasing competition among agricultural, municipal, industrial, and
ecological uses. Although water management practices in the United States are generally advanced,
particularly in the West, the reliance on past conditions as the basis for current and future planning may
no longer be appropriate, as climate change increasingly creates conditions well outside of historical
observations. Rising temperatures will diminish snowpack and increase evaporation, affecting seasonal
availability of water. In the Great Lakes and major river systems, lower water levels are likely to
exacerbate challenges relating to water quality, navigation, recreation, hydropower generation, water
transfers, and binational relationships. Decreased water supply and lower water levels are likely to
exacerbate challenges relating to aquatic navigation in the United States.
Higher water temperatures, increased precipitation intensity, and longer periods of low flows will
exacerbate many forms of water pollution, potentially making attainment of water quality goals more
difficult. As waters become warmer, the aquatic life they now support will be replaced by other species
better adapted to warmer water. In the long term, warmer water and changing flow may result in
deterioration of aquatic ecosystems.
Climate change is likely to affect U.S. energy use and energy production and physical and
institutional infrastructures. It will also likely interact with and possibly exacerbate ongoing
environmental change and environmental pressures in settlements, particularly in Alaska where
indigenous communities are facing major environmental and cultural impacts. The U.S. energy sector,
which relies heavily on water for hydropower and cooling capacity, may be adversely impacted by
changes to water supply and quality in reservoirs and other water bodies. Water infrastructure, including
drinking water and wastewater treatment plants, and sewer and stormwater management systems, will be
at greater risk of flooding, sea level rise and storm surge, low flows, and other factors that could impair
performance.
Disturbances such as wildfires and insect outbreaks are increasing in the United States and are
likely to intensify in a warmer future with warmer winters, drier soils, and longer growing seasons.
Although recent climate trends have increased vegetation growth, continuing increases in disturbances are
likely to limit carbon storage, facilitate invasive species, and disrupt ecosystem services.
Over the 21st century, changes in climate will cause species to shift north and to higher elevations
and fundamentally rearrange U.S. ecosystems. Differential capacities for range shifts and constraints
from development, habitat fragmentation, invasive species, and broken ecological connections will alter
ecosystem structure, function, and services.
Climate change impacts will vary in nature and magnitude across different regions of the United
States.
Prepublication Copy
Appendix C 121
• Sustained high summer temperatures, heat waves, and declining air quality are projected in
the Northeast, 3 Southeast, 4 Southwest, 5 and Midwest. 6 Projected climate change would
continue to cause loss of sea ice, glacier retreat, permafrost thawing, and coastal erosion in
Alaska.
• Reduced snowpack, earlier spring snowmelt, and increased likelihood of seasonal summer
droughts are projected in the Northeast, Northwest, 7 and Alaska. More severe, sustained
droughts and water scarcity are projected in the Southeast, Great Plains, 8 and Southwest.
• The Southeast, Midwest, and Northwest in particular are expected to be impacted by an
increased frequency of heavy downpours and greater flood risk.
• Ecosystems of the Southeast, Midwest, Great Plains, Southwest, Northwest, and Alaska
are expected to experience altered distribution of native species (including local extinctions),
more frequent and intense wildfires, and an increase in insect pest outbreaks and invasive
species.
• Sea level rise is expected to increase storm surge height and strength, flooding, erosion, and
wetland loss along the coasts, particularly in the Northeast, Southeast, and islands.
• Warmer water temperatures and ocean acidification are expected to degrade important
aquatic resources of islands and coasts such as coral reefs and fisheries.
• A longer growing season, low levels of warming, and fertilization effects of carbon dioxide
may benefit certain crop species and forests, particularly in the Northeast and Alaska.
Projected summer rainfall increases in the Pacific islands may augment limited freshwater
supplies. Cold-related mortality is projected to decrease, especially in the Southeast. In the
Midwest in particular, heating oil demand and snow-related traffic accidents are expected to
decrease.
Climate change impacts in certain regions of the world may exacerbate problems that raise
humanitarian, trade, and national security issues for the United States. The IPCC identifies the most
vulnerable world regions as the Arctic, because of the effects of high rates of projected warming on
natural systems; Africa, especially the sub-Saharan region, because of current low adaptive capacity as
well as climate change; small islands, due to high exposure of population and infrastructure to risk of sea
level rise and increased storm surge; and Asian mega-deltas, such as the Ganges-Brahmaputra and the
Zhujiang, due to large populations and high exposure to sea level rise, storm surge and river flooding.
Climate change has been described as a potential threat multiplier with regard to national security issues.
3
Northeast includes West Virginia, Maryland, Delaware, Pennsylvania, New Jersey, New York, Connecticut, Rhode
Island, Massachusetts, Vermont, New Hampshire, and Maine.
4
Southeast includes Kentucky, Virginia, Arkansas, Tennessee, North Carolina, South Carolina, southeast Texas,
Louisiana, Mississippi, Alabama, Georgia, and Florida.
5
Southwest includes California, Nevada, Utah, western Colorado, Arizona, New Mexico (except the extreme
eastern section), and southwest Texas.
6
The Midwest includes Minnesota, Wisconsin, Michigan, Iowa, Illinois, Indiana, Ohio, and Missouri.
7
The Northwest includes Washington, Idaho, western Montana, and Oregon.
8
The Great Plains includes central and eastern Montana, North Dakota, South Dakota, Wyoming, Nebraska, eastern
Colorado, Nebraska, Kansas, extreme eastern New Mexico, central Texas, and Oklahoma.
Prepublication Copy