From Hurricanes to Epidemics: The Ocean's Evolving Impact on Human Health - Perspectives from the U.S. (Global Perspectives on Health Geography) 3030550117, 9783030550110

This book provides a complete and current overview of the correlation between ocean conditions and human health, publish

115 66 5MB

English Pages 201 [193] Year 2020

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Preface
Acknowledgments
Contents
Contributors
About the Author
List of Figures
List of Tables
Chapter 1: Introduction
1.1 Changing Ocean Currents
1.2 Evolving Patterns of Marine Food Chains
1.3 The Blue Mind: The Science Behind the Mental Health Benefits of Aquatic Environments
1.4 Marine Plastic Contamination
1.5 COVID-19 and the Oceans: Impacts on the Environment, the Blue Economy, and Ocean Health
1.6 Historical Impact of Epidemics on the Environment
1.7 The Blue Economy
1.8 Challenges to Operational Ocean Science
References
Chapter 2: Changing Ocean Currents
2.1 Introduction
2.2 Influence of Changing Ocean Currents on Spatial Patterns of Sea Level
2.3 Shifting Ocean Currents and Predicted Influence on Future Climate
2.4 The Role of Ocean Circulation in Marine Heat Waves, Acidification, and Deoxygenation
2.5 Conclusions
References
Chapter 3: Health Consequences of Marine Oil Spills: Lessons Learned from the Deepwater Horizon Accident
3.1 Offshore Oil Exploration
3.2 Source of Marine Oil
3.3 Composition of Marine Oil
3.4 Human Toxicity
3.5 Epidemiological Studies on Oil Spills
3.6 Deepwater Horizon: Lessons Learned
3.7 Ecosystem Effects of Deepwater Horizon Oil Spill
3.8 Acute Health Impact
3.9 Long-Term Health Consequences of Oil Spills
3.10 Mental Health Outcomes: Association with Health Disparities
3.11 Risk Paradigm
3.12 Impact on Woman and Children
3.13 Conclusion
Resources
Chapter 4: New England Coastal Cities: The Struggle for a Resilient Future
4.1 New England Coastal Cities: The Struggle for a Resilient Future
4.2 Economic Struggles
4.3 Sea Level Rise and Extreme Rain Events
4.4 Developing Strategies for Resilience
4.5 The Boston Example
4.6 Moving Resilience Forward in Southern New England Cities
4.6.1 New Bedford
4.6.2 Bridgeport
4.6.3 Providence
4.7 Resilience Building in Southern New England Cities: Summary and Questions
References
Chapter 5: The Blue Mind
5.1 The Blue Gym: Threats to the Mental Health Benefits of the Ocean
5.2 The Blue Gym: Research Outcomes
5.3 Threats to the Blue Gym
5.4 The Need to Communicate Success Stories
References
Chapter 6: Oceans and Rapid Climate Change
6.1 Insights from Past Climates
6.2 The Stability of the AMOC
6.3 Future AMOC Variations and Climate Changes
References
Chapter 7: Evolving Marine Health Threats to Humans
7.1 Rising Sea Temperatures and the Spread of Pathogens: The Example of Vibrio vulnificus Infections
7.2 Disease Outbreak: Correlation with Sea Temperature and Climate Change
7.3 Melting Permafrost
7.4 Harmful Algal Blooms (HAB): Cyanobacteria and Red Tides
7.5 Cyanobacteria Lake, Pond, and River: Blue-Green Algal Blooms
7.6 Red Tides: Costal Blooms
7.6.1 Long-Term Impact of HAB on Human Health
7.6.2 Economic Impact of HAB
7.6.3 Conclusion: Outlook for HAB and Their Effect on Human Health
7.6.4 Decreasing Biodiversity: Threats to the Blue Biotechnology Industry
7.6.5 Ocean Currents and Disease Modeling
References
Chapter 8: International Maritime Law and its Applications for the Twenty-First Century
Bibliography
Chapter 9: Marine Pharmacology: Threats to Undiscovered Sources of Medical Therapeutics
9.1 Decreasing Biodiversity: Threats to the Blue Biotechnology Industry
References
Chapter 10: Tropical Cyclones
References
Chapter 11: Microplastic Invasion – A Threat to Animal and Human Health
11.1 Plastic Crisis
11.2 What Are Macro-, Micro-, and Nanoplastics?
11.3 How Do Microplastics Make It to the Ocean?
11.4 Breakdown of Plastic into Microplastics
11.5 Areas in the Ocean Where Plastic Is Most Prevelant
11.6 Effects on the Marine Life
11.7 Human Ingestion
11.8 Human Health Effects
11.9 Socioeconomic Factors
11.10 Neurodevelopmental and Psychological Effects
11.11 Nanoplastics
11.12 Conclusion
Bibliography
Chapter 12: The Oceans and the COVID-19 Pandemic: Explaining the Relationship with the Systems View of Life
12.1 Introduction
12.2 The Systems View of Life
12.3 Complex Adaptive Systems
12.4 The Ocean Is a Complex Adaptive System
12.5 The Looming Environmental Crisis Ahead of Us
12.6 Practical Application of Systems Thinking
12.7 COVID-19, Humanity, and the Ocean
12.8 Conclusion
References
Chapter 13: Oceans and Human Health: A Rising Tide of Challenges and Opportunities
13.1 Oceans and Human Health: A Rising Tide of Challenges and Opportunities
13.2 United Nations Sustainable Development Goal 14: Conserve and Sustainably Use the Oceans, Seas, and Marine Resources
13.3 Sustainable Use of the Oceans, Seas, and Marine Resources: Success Stories
13.4 The Blue New Deal
References
Index
Recommend Papers

From Hurricanes to Epidemics: The Ocean's Evolving Impact on Human Health - Perspectives from the U.S. (Global Perspectives on Health Geography)
 3030550117, 9783030550110

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

Global Perspectives on Health Geography

Kevin Michael Conrad Editor

From Hurricanes to Epidemics The Ocean’s Evolving Impact on Human Health - Perspectives from the U.S.

Global Perspectives on Health Geography Series Editor Valorie Crooks, Department of Geography, Simon Fraser University, Burnaby, BC, Canada

Global Perspectives on Health Geography showcases cutting-edge health geography research that addresses pressing, contemporary aspects of the health-place interface. The bi-directional influence between health and place has been acknowledged for centuries, and understanding traditional and contemporary aspects of this connection is at the core of the discipline of health geography. Health geographers, for example, have: shown the complex ways in which places influence and directly impact our health; documented how and why we seek specific spaces to improve our wellbeing; and revealed how policies and practices across multiple scales affect health care delivery and receipt. The series publishes a comprehensive portfolio of monographs and edited volumes that document the latest research in this important discipline. Proposals are accepted across a broad and ever-developing swath of topics as diverse as the discipline of health geography itself, including transnational health mobilities, experiential accounts of health and wellbeing, global-local health policies and practices, mHealth, environmental health (in)equity, theoretical approaches, and emerging spatial technologies as they relate to health and health services. Volumes in this series draw forth new methods, ways of thinking, and approaches to examining spatial and place-based aspects of health and health care across scales. They also weave together connections between health geography and other health and social science disciplines, and in doing so highlight the importance of spatial thinking. Dr. Valorie Crooks (Simon Fraser University, [email protected]) is the Series Editor of Global Perspectives on Health Geography. An author/editor questionnaire and book proposal form can be obtained from Publishing Editor Zachary Romano (zachary. [email protected]).

More information about this series at http://www.springer.com/series/15801

Kevin Conrad Editor

From Hurricanes to Epidemics The Ocean’s Evolving Impact on Human Health - Perspectives from the U.S.

Editor Kevin Conrad Ochsner Health New Orleans, LA, USA

ISSN 2522-8005     ISSN 2522-8013 (electronic) Global Perspectives on Health Geography ISBN 978-3-030-55011-0    ISBN 978-3-030-55012-7 (eBook) https://doi.org/10.1007/978-3-030-55012-7 © Springer Nature Switzerland AG 2021 This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. The publisher, the authors, and the editors are safe to assume that the advice and information in this book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty, expressed or implied, with respect to the material contained herein or for any errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. This Springer imprint is published by the registered company Springer Nature Switzerland AG The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland

To the resilient people of coastal Louisiana. Survivors of hurricanes, wars, and epidemics, they now face their greatest challenge from the rising seas, land loss, and a rapidly changing ocean environment.

Preface

It is an interesting biological fact that all of us have, in our veins the exact same percentage of salt in our blood that exists in the ocean, and, therefore, we have salt in our blood, in our sweat, in our tears. We are tied to the ocean. And when we go back to the sea, whether it is to sail or to watch it we are going back from whence we came.—John F. Kennedy

I have been a part of three great health challenges in my medical career: first was the AIDS epidemic, then the Opioid epidemic, and now the COVID-19 pandemic. Despite the tremendous loss of life and suffering in each of these, perhaps our greatest challenge will be the climate crisis and its effect on human health. There are similarities in each of these. In each instance, the issues were in part man made and the problems seem unsolvable. In each circumstance lack of inertia seem to be the greatest obstacle. At that time and now, how to get started with such a massive problem seems to be the greatest challenge. We are currently and rightly focused on our most immediate threat, the COVID-19 pandemic, but issues such as the declining health of our environments and oceans have not gone away. Although the AIDS epidemic, the opioid epidemic, and the current COVID-19 pandemic is not solved by any calculation, we have made progress. First by leaders who sounded the alarms and then by those who took action despite any visible endpoint. It took taking the problems and breaking them down into facts, which demonstrated that issues had a direct impact on our daily lives, and taking small bites. As the proverb states, “How do you eat an elephant? One bite at time.” I think key to this process is finding a common voice that speaks to all people. When focusing on public health, why single out the oceans? We recognize the ocean as a distinct body, albeit one that is very large. One that through literature and the arts can be identified with on a personal level. The ocean’s health and functioning is a reliable marker for the overall health of the planet. Few would argue that a healthy ocean is not essential for our future. Climate change may be an abstract concept to many, but intuitively we know the risks that are posed by an imbalanced ocean. It is impossible to ignore an entire blanched coral reef, an expanding dead zone, or a beach cluttered with plastic. Maybe this perception and visual imprint will drive humankind towards urgently needed action. vii

viii

Preface

Increasingly, it has been established that the health of the ocean is intimately tied to our health. Waterborne infectious diseases, harmful algal bloom toxins, contaminated seafood, and chemical pollutants are signals that ocean imbalance is placing humankind in jeopardy. Just as we have threatened the health of our ocean, the ocean too can threaten our health. Our coastal economies are threatened as well. Communities that have once thrived on the seafood industry struggle to survive. Economic hardships in these regions are leading to declining physical and mental health. The oceans systems are highly interconnected and these relations are being understood only now. Physical change to these systems is occurring at unprecedented rates, each feeding upon the other. Predictions made, such as sea rise, are often inaccurate without first fully understanding the impact of other forces. As with many great challenges, understanding the impact of the oceans on human health will take the efforts and voices form a variety of fields. To this end, I have attempted to include chapters from a variety of disciplines. Some at first glance have no bearing on human health, but with deeper understanding, the connections emerge. Certainly, the COVID-19 pandemic will allow us to take a different view of our planet. One where an understanding of those connections is not just an academic exercise but essential for our survival. Solutions will come from international cooperation and regulations. The chapter on maritime law explores the complexities of enforcing environmental regulations across countries with variable economic development. Fortunately, countries with different levels of development are expressing a commitment to the health of our oceans. The significance of the United Nations cannot be underemphasized. This has been a collaborative work by many authors from various fields. I am appreciative of their efforts and dedication to their respective fields. As with many, their current research is being challenged by circumstances. I would like to specially thank Christy Mo, my managing editor, for her efforts in identifying and recruiting the contributors. New Orleans, LA, USA  Kevin Conrad

Acknowledgments

I would like to thank Christy Mo, my managing editor, in the coordination and editing of this text. Her combined interest in the interaction of the environment and human health was invaluable in the preparation of this text.

ix

Contents

1 Introduction����������������������������������������������������������������������������������������������    1 Kevin Conrad 2 Changing Ocean Currents����������������������������������������������������������������������   11 David Trossman and Jaime Palter 3 Health Consequences of Marine Oil Spills: Lessons Learned from the Deepwater Horizon Accident��������������������������������������������������   27 Kevin Conrad, Rea Cleland, and Nicholas Reyes 4 New England Coastal Cities: The Struggle for a Resilient Future������   39 H. Curtis Spalding, Siddhi Nadkarni, Claire Bekker, and Devyn Collado-Nicol 5 The Blue Mind������������������������������������������������������������������������������������������   59 Kevin Conrad, Rea Cleland, and Nicholas Reyes 6 Oceans and Rapid Climate Change�������������������������������������������������������   67 Wei Liu and Alexey Fedorov 7 Evolving Marine Health Threats to Humans����������������������������������������   81 Kevin Conrad, Rea Cleland, and Nicholas Reyes 8 International Maritime Law and its Applications for the Twenty-First Century������������������������������������������������������������������   95 Swati Parashar 9 Marine Pharmacology: Threats to Undiscovered Sources of Medical Therapeutics��������������������������������������������������������������������������  115 Rea Cleland and Kevin Conrad 10 Tropical Cyclones������������������������������������������������������������������������������������  121 Isaac Ginis

xi

xii

Contents

11 Microplastic Invasion – A Threat to Animal and Human Health��������  129 Joseph Casper III 12 The Oceans and the COVID-19 Pandemic: Explaining the Relationship with the Systems View of Life������������������������������������  149 Marianne Maumus 13 Oceans and Human Health: A Rising Tide of Challenges and Opportunities������������������������������������������������������������������������������������  167 Kevin Conrad Index������������������������������������������������������������������������������������������������������������������  173

Contributors

Claire Bekker  Brown University, Geology-Biology, Providence, RI, USA Joseph Casper III  University of New Orleans, New Orleans, LA, USA Rea  Cleland  University of Queensland-Ochsner Clinical School, New Orleans, LA, USA Devyn Collado-Nicol  Brown University, Environment and Enterprise, Providence, RI, USA Kevin Conrad  Ochsner Health, New Orleans, LA, USA Alexey  Fedorov  Department of Geology and Geophysics, Yale University, New Haven, CT, USA LOCEAN/IPSL, Sorbonne University, Paris, France Isaac Ginis  University of Rhode Island, Kingston, RI, USA Wei  Liu  Department of Earth and Planetary Sciences, University of California Riverside, Riverside, CA, USA Marianne Maumus, MD  Ochsner Health Systems, New Orleans, LA, USA Siddhi Nadkarni  Brown University, Health and Human Biology, Providence, RI, USA Jaime  Palter  Graduate School of Oceanography, University of Rhode Island, Kingston, RI, USA Swati Parashar  Tulane University School of Law, New Orleans, LA, USA Nicholas Reyes  University of Queensland-Ochsner Clinical School, New Orleans, LA, USA H.  Curtis  Spalding  Institute at Brown University for Environment and Society, Providence, RI, USA David Trossman  University of Texas at Austin, Oden Institute for Computational Engineering and Sciences, Austin, TX, USA xiii

About the Author

Kevin  Conrad  is a practicing physician with Ochsner Health in New Orleans, Louisiana. He is an associate professor of Medicine at Tulane Medical school and as senior instructor at the Queensland school of medicine in Australia. He serves as Medical Director of Community Affairs at Ochsner Health. He has been active in promoting green initiatives within the healthcare system and has written and lectured on the topic nationally. His research interests include social and environmental impacts on health.

xv

List of Figures

Fig. 2.1 At the center is a map of the rate of change in sea surface height between 1993 and 2012 from satellite altimetry. Also shown are relative sea level changes (gray lines) for various tide gauges for the period 1995–2012. For comparison, an estimate of global mean sea level change is shown in each panel (red line). (Reproduced from the Intergovernmental Panel on Climate Change Fifth Assessment Report Chap. 13 on Sea Level Rise [97]).......�������������������������������������������������� 13 Fig. 2.2 Schematic of the Atlantic Meridional Overturning Circulation (top) and the modeled projection of surface air temperature response to an idealized global warming simulation (bottom). The schematic shows the Gulf Stream and its northeastward extension, the North Atlantic Current (NAC) transporting warm water to the Nordic and Labrador Seas where they cool and are transformed to dense water masses that flow near the seafloor. The Deep Western Boundary Current (DWBC) transports the cold, dense waters equatorward. The moorings used to monitor the AMOC strength are depicted in red. This schematic was reproduced from [33]. Panels b–d are from model simulations in which atmospheric CO2 is doubled at a rate of 1% per year. Panel b shows surface air temperature change in response to the CO2 increase in a normal simulation, in which ocean circulation responds dynamically to the warming. Panel c shows the surface temperature change after a doubling of CO2 in a simulation in which the velocity field is replaced with that from a control simulation (i.e., with no change in atmospheric CO2). Panel c shows the zonal average surface air temperature response in the normal model (black) and the model with fixed ocean circulation (blue). The difference between the maps and zonal averages reveals that the change in the large-scale circulation (including a 25% slowdown of the AMOC) slows the pace of global warming, principally due to cooling at high northern latitudes. (Maps and zonal averages reproduced from [39])���������������������������������������������������������������������������������������������� 15

xvii

xviii

List of Figures

Fig. 2.3 A schematic of the stratospheric and tropospheric polar vortex, the latter of which is more commonly known as the jet stream. Meanders of these vortices can produce anomalous weather at midlatitudes. An active area of debate is how these dynamical features will respond to warming, with evidence mounting for a wavier jet stream and altered polar vortices. (Modified from [98])���������������������������������� 17 Fig. 2.4 Future biogeochemical change in the world’s oceans. Maps a–c show the difference between future (i.e., the average from 2091 to 2100) and contemporary (i.e., the average from years 1996 to 2005) values in a future in which CO2 emissions are not reduced under any policy interventions (called RCP8.5). Plots d–f show the global average change relative to contemporary values under RCP8.5 and a future with moderate emission reductions (called RCP4.5) at the ocean surface and seafloor. (Modified from [99]) ������������������������������������������ 18 Fig. 3.1 Risk paradigm �������������������������������������������������������������������������������������� 35 Fig. 4.1 Total population decline across Providence, New Bedford, and Bridgeport�������������������������������������������������������������������������������������� 42 Fig. 4.2 Percentage in poverty in Bristol County, MA [10] ������������������������������ 43 Fig. 4.3 Global sea level rise scenarios for the United States National Climate Assessment (2018) United States Department of Commerce. https:// nca2018.globalchange.gov/chapter/2/. Accessed 27 July 2020������������ 45 Fig. 4.4 Observed change in very heavy precipitation. (Taken from United States National Climate Assessment (2014) United States Department of Commerce. https://nca2014.globalchange.gov/ report/our-changing-climate/heavy-downpours-increasing. Accessed 27 July 2020)������������������������������������������������������������������������ 46 Fig. 4.5 Major vulnerability factors to climate change [42]������������������������������ 54 Fig. 6.1 Atlantic meridional overturning stream function (units: Sverdrup or Sv = 106 m3/s) based on the annual mean data of the ECCO (Estimating the Circulation and Climate of the Ocean) dataset during 1992–2015 [73]. Black arrows indicate the circulation direction of the AMOC ������������������������������������������������ 68 Fig. 6.2 A temperature-based AMOC index (blue) reconstructed from the NASA GISS surface temperature data [74] based on the method by Ramhstorf et al. [75], and the AMOC strength from the RAPID-MOCHA observations (red, c.f. [2])���������������������������������� 68 Fig. 6.3 Evidence for a weaker AMOC during Heinrich and the Younger Drays events (marked as gray bands). Oxygen isotope ratio in the planktonic foraminifer Neogloboquadrina pachyderma from the western North Atlantic [76] is shown in green; low values reflect the presence of glacial meltwater. 231Pa/230Th ratio measured in sediments on the Bermuda Rise is shown in purple [29], plum [77], and magenta [78], which reflects changes in the residence time of deep water in the

List of Figures

xix

Atlantic potentially related to changes in the AMOC—higher values mean a weaker AMOC. Note that the vertical axis is inverted for the top and middles curves. Mean 6C13 values for the deep Atlantic (an average between several records) are shown in blue [28]; low values indicate the prevalence of southern source waters, rather than northern source water, which may indicate AMOC weakening. Additional abbreviations: kyr BP 1000 years before present; HS Heinrich stadial; YD Younger Dryas. Note that there are uncertainties on the dating of these events in different time series������������������������������������������������������ 70 Fig. 6.4 (a) Variations in the AMOC strength (estimated as the maximum stream function below 500 m in the North Atlantic) in the original (blue) and corrected (red) configurations of the NCAR CCSM3 [62] for the 1990 carbon dioxide level (prior to the year 200) and after atmospheric carbon dioxide concentration doubling (after the year 200). Annual mean surface temperature change (in K) 300 years after atmospheric carbon dioxide concentration has been doubled in the (b) original and (c) corrected models���������������������������� 75 Fig. 7.1 Annual global sea surface temperature anomalies from 1880 to 2015 with superimposed linear trend. (From: http://www.ncdc.noaa. gov/cag/time-series/global/globe/ocean/ytd/12/1880-2016)���������������� 85 Fig. 8.1 From: https://lawexplores.com/imo-intsitutional-structure-and-lawmaking-process/������������������������������������������������������������������������������������ 96 Fig. 8.2 From: https://news.un.org/en/story/2015/12/519172-sustainable-development-goals-kick-start-new-year�������������������������������������������������������� 97 Fig. 8.3 From: https://theoceancleanup.com/great-pacific-garbage-patch/�������� 98 Fig. 8.4 Map of parties to the London convention/protocol. Legend: Green: Protocol parties, Yellow: convention parties, Red: non-parties. Status on 22 February 2019. (From: http://www.imo.org/en/OurWork/ Environment/LCLP/Documents/Parties%20to%20the%20LCLP%20 February%202019.pdf) ���������������������������������������������������������������������� 100 Fig. 8.5 From:  http://www.imo.org/en/OurWork/Environment/LCLP/TC/ Documents/London%20Protocol%20Why%20it%20is%20needed%20 20%20years.pdf���������������������������������������������������������������������������������� 101 Fig. 8.6 The carbon cycle. (From: https://cnx.org/contents/s8Hh0oOc@ 9.10 :1KV9fus6@4/Biogeochemical-Cycles)������������������������������������ 103

List of Tables

Table 4.1 Demographic indicators in Providence, New Bedford, and Bridgeport vs. national average (Census Bureau)�������������������������� 43 Table 7.1 List of common waterborne diseases���������������������������������������������������� 83 Table 7.2 Pharmaceuticals derived from marine sources�������������������������������������� 90 Table 9.1 Pharmaceuticals derived from marine sources������������������������������������ 118

xxi

Chapter 1

Introduction Kevin Conrad

When we try to pick out anything by itself, we find it hitched to everything else in the Universe. —John Muir

Our health, well-being, and survival depend on the health of the oceans and the oceans are not well. In the past four decades, plastic pollution in the sea has increased tenfold, a third of fish stocks are now threatened, deoxygenated dead zones are expanding, and ocean acidification endangers a multitude of species. Sea level rise and other impacts of climate change now threaten multiple coastal areas where much of our population resides. Floodwaters and catastrophic climate events are now a significant cause of human displacement, economic turmoil, and human migration. These changes, and in particular events that have resulted in environmental refugees, have now endangered political stability globally. The state of the oceans is having a direct impact on human well-being, and it is imperative that we understand the intertwined links of ocean ecology and how we can take action to preserve this delicate relationship. Interconnected ocean systems such as currents, chemistries, ice formation, and temperature are undergoing rapid change. Each system feeds off each other and produces rapid exponential change resulting in unexpected weather events, unusual algal blooms, and coastal habitat loss. Marine system connections are being connected at unprecedented rate, and temperature rise is leading to ice melts which in turn is affecting currents that redistribute nutrient loads. All life-forms, including humankind, are ill equipped to deal with these rapid oscillations. Life is adaptable to a degree, but for the most part, prospers with relative environmental stability which has historically been provided by the earth’s oceans. Certainly, the rate of change seen in the oceans may be far too much for many life-forms that have existed

K. Conrad (*) Ochsner Health, New Orleans, LA, USA e-mail: [email protected] © Springer Nature Switzerland AG 2021 K. Conrad (ed.), From Hurricanes to Epidemics, Global Perspectives on Health Geography, https://doi.org/10.1007/978-3-030-55012-7_1

1

2

K. Conrad

for millennia to sustain, and some have suggested that we are entering a humaninduced mass marine extinction event [1]. With each new report released, change seems to be accelerating. The evidence that the oceans are declining as a viable ecosystem is constant, predictable, and exponentially growing. Marine scientists are some of the world’s most tireless and dedicated explorers. They have been sounding the alarms for decades, but now their voice has a sense of urgency. Over the past several years, scientists have developed new technologies, explored new depths, and made new discoveries. With each new area explored comes alarming evidence of the human impact on the ocean environment. Each day brings new revelations of extensive algal blooms, unexpected fish kills, dead zones, melting glaciers, and massive coral reef destruction. Research reveals that through the years of filling the oceans with plastic product, the debris has not only posed a physical barrier to sea life but has infused all aquatic life with microplastics, whose impact on the health of humans is not yet known. A systems view of life may provide us with the best explanation of these connections. If the earth is viewed as a single organism, then the oceans represent as its most essential organ. It is responsible for oxygenation, circulation, digestion, and detoxification. In life, organs under stress tend to decline in exponential fashion rather than in a direct line. An alcoholic, after years of drinking, will suddenly experience liver failure. A person with hypertension may suddenly develop kidney failure. Coronary artery disease may present with a sudden heart attack. One of the challenges of treating these conditions is convincing patients that damage is occurring despite overt obvious signs. One organ failure often leads to another. Organs tend to decline, suddenly and irreversibly. Are the oceans at that threshold? Are we seeing signs of imminent failure? Marine biologists, oceanographers, and environmentalists have been documenting the rapid change in ocean environments for decades. Their constant warnings have reminded us the far-ranging impact that humans can have. These alarms have taken a different tone recently. The alarms have shifted toward a distinct timetable that mankind has before irreversible damage occurs. As scientists, it is their responsibility not only to discover the facts but to connect the dots and communicate these facts effectively and repeatedly. In a world saturated with a constant new cycle, this has proved to be a challenge. Recently, there have been constant visible reminders that this vast ecosystem once thought to be beyond the effects of human activity has reached a tipping point, one that many pose an immediate threat to man’s survival. Many worrisome environmental topics occupy the news cycle, but the greatest, the one that ties all factors together, may be man’s interaction and dependency with the oceans. Civilization has evolved in direct correlation with the flow of the seas. The ocean has always shaped human history in profound and unexpected ways. The sea has provided trade routes, fueled empires, and enriched the arts. With its powerful effect on the planet’s climate, the sea has influenced human progress and today continues to affect us wherever we dwell, whether in mountains, deserts, or cities. It is logical, if not somehow overlooked, that drastic changes on the oceans will bring about profound effects on society.

1 Introduction

3

1.1  Changing Ocean Currents Ocean currents have always been a major influence on civilizations’ progress. The currents determined where trade routes were established and where civilizations were developed. Changes in currents have had a marked influence on regions, and there is evidence that these currents are changing. These currents have at times influenced prosperity and with alterations have brought about disaster. Throughout history, there are many examples of the profound effect of ocean currents. The Yucatan Peninsula has water that supplied seasonally, by monsoon rains blowing in from the sea. Changes in Atlantic winds, currents, and sea surface temperatures have caused multiyear droughts. The Mayan culture existed over 300  years ago. Around 750–900  AD, it underwent a sudden and catstrophic collapse. This is thought to be due to changes in climates and currents affecting the region. Due to these changes, Mayan reservoirs and canals could no longer support its maize-based agriculture leading to eventual decline [2]. It is foolhardy to think that modern societies aren’t susceptible to similar events. Oceans have attracted humans to their shores and waters since the ancient time. Over thousands of years, they have served as a source of food, provided livelihoods, and generated commerce. They have connected peoples, molded civilizations, and defined who we are as a society. Their importance is reflected in art, music, architecture, and our cultural practices. Human development and survival are permanently linked to ocean habitats.

1.2  Evolving Patterns of Marine Food Chains Economies dependent on ocean habitats both as food source and tourism are now being affected. In the Caribbean, where many islands are dependent on tourism, massive blooms of Sargasso seaweed (Sargassum) over a meter thick have washed up on the shore releasing irritating and toxic sulfur compounds [3]. This has had a negative impact on tourism. Before 2011, open-ocean Sargassum was mostly confined to the Sargasso Sea. This was due to an enclosed North Atlantic ocean current. This area also served as a spawning ground for eels. Originally, it was assumed that the Sargassum drifted south from the Sargasso Sea. Satellite imagery and data on ocean currents revealed that new local blooms and changing currents were to blame. Since 2011, blooms have occurred yearly. Algal blooms occur globally. In Chile, a massive bloom of Alexandrium catenella contaminated an offshore salmon farm killing over 100,000 tons of salmon [4]. This devastated their aquaculture industry and led to protests. Algal blooms continue to plague the Florida and other southern states in the United States. Animal studies have shown that long-term exposure to these toxins can result in harm to developing brains [5]. The effects of these neurotoxins on human development have yet to be elucidated.

4

K. Conrad

The deepest and most remote places on earth may hold the answers to many of our greatest medical challenges. Seventy-five percent of the earth’s surface is covered by water, but research into the pharmacology of marine organisms has been limited. Oceans contain more than 80% of the plant and animal species in the world, yet practically, all pharmaceuticals have been derived from easily obtained land-­ based organisms. The potential of the oceans to provide new medical life-changing discoveries remains untapped [6]. The marine environment is an exceptional storehouse of unique bioactive products. Evolving in isolation over many centuries, marine life contains structural and chemical features generally not found in land life-forms. Marine animals survive under extreme bariatric pressures, excessive cold, and hot temperatures and withstand toxic chemical exposures which would be lethal to all terrestrial life. Under these conditions, many ocean species have developed unique molecular adaptations to survive these harsh environments. The potential of these unique molecular products is yet to be explored and unless preserved will be lost to science.

1.3  T  he Blue Mind: The Science Behind the Mental Health Benefits of Aquatic Environments There is increasing evidence that being near or on the water has a significant impact not only on our mental but also on physical health. The chapter on the blue mind explores this concept and reports on the research that supports it. The mere act of providing good stewardship of our ocean ecosystems seems to play an important role in this benefit [7]. Our interaction with the oceans has been on some levels straightforward such as a source of food, enjoyment, and its profound effect on the earth’s meteorological events. On other levels, it is more complex, such as the effect of CO2 production, the emergence of new infectious diseases, the benefits of marine biopharmaceuticals, and its contributions to biodiversity. In each of these areas, human activity has had a marked influence on the ocean’s biology, chemistry, and currents. This effect was thought to be manageable but new research suggest otherwise. How changes in the ocean are impacting human mental health on a global scale is recently being explored. Weather we reside inland or in coastal region, the oceans impact each of us profoundly. Now more than ever, the mental benefits of the blue mind may be needed.

1.4  Marine Plastic Contamination For decades, marine vessels and onshore activities have filled oceans with plastic materials. Even in the most remote areas, such as the Marianas Trench, plastic material can be found. For decades, there were few regulations or governing bodies that

1 Introduction

5

limit plastic pollution. Finally, in 1972, the Convention on the Prevention of Marine Pollution by Dumping of Wastes and Other Matter, known as the London Convention or MARPOL, was ratified. Further regulations were further introduced by the London Protocol, an update to the convention that began in 2006. Despite these accords, plastic continues to accumulate in the oceans at an alarming rate, and despite the apparent vastness of the oceans, waste is visibly present in most regions. Chapter 11 examines the source of the plastics and how treaties and regulations can be utilized to combat this issue and what framework should be adopted to address this issue.

1.5  C  OVID-19 and the Oceans: Impacts on the Environment, the Blue Economy, and Ocean Health In a matter of months, the world and our perception of the world has been transformed. A pandemic that began in China now has become an immediate and existential threat to our existence. The loss of life has been unimaginable and our future seems unpredictable. Epidemics have occurred before, but we have relied and expected our experts to manage and limit their impacts and have had little impact to the majority of the world’s population. For the most part, they have been geographically limited in their impact. The world now seems a much smaller and much more vulnerable place. Will this crisis bring to people’s minds that maybe the risk scientists have been talking about with climate change is not so far-fetched? Chapter 12 written by Marianne Maumus explores how the pandemic will hopefully lead to a systems view of the planet and our marine environment. Hopefully, a common interconnected view of marine systems by both scientists and nonscientists will lead to the adoption of a common voice. The pandemic has brought about widespread job losses, stalled economic activity, and severely restricted all forms of transportation. With this have come drops in carbon emissions. In China, emissions fell by 25% in the beginning of the year 2020, and in New York, pollution levels have declined by 50% [8, 9]. Similar results have been reported from around the world. Landmarks, such as the Himalayas that have been hidden by haze for decades in the Indian state of Punjab, suddenly have become visible. The canals of Venice are clearer than anyone can remember, and Mount Kenya is visible in Nairobi [10]. A view of the world with cleaner air and water has been presented to many people for the first time. To many who seem skeptical of our ability to impact the environment, this is a firsthand evidence that a different world is possible. The change in air quality is expected, as transportation, which has been severely restricted, make up 28.2% of carbon emissions. Driving makes up 72% of the total and aviation contributes 11% [11]. Unnecessary travel has been severely restricted and long distance trips have been postponed.

6

K. Conrad

Even in the short term, the pandemic will not be enough to address the reduction in carbon emissions needed to effect climate change. A report from the Carbon Brief states the pandemic may lead to climate-warming greenhouse gas emissions dropping by 5.5% in 2020 as compared to the last year [12]. Although this would be a historic drop, it is only two thirds of what is needed to keep warming to 1.5 C above pre-industrial levels. Global emissions would need to fall by 7.6% each year this decade, in order to limit warming to less than 1.5 C above pre-industrial temperatures. This is the goal set by the United Nations in their 2018 consensus report [13]. Will these changes persist? There is no certain end to the restrictions imposed, so in the short term, the answer is probably yes. In the long term, it is more difficult to predict. When the economy opens up, as expected in a graduated fashion, several factors may contribute to an overall rebound that results in either an increase or decrease in carbon emissions. Certainly, when self-isolation is relaxed, people will be inclined to travel more. Will we have adapted to a new lifestyle that has learned to do and enjoy more activities from home? Business that once had the economic ability to travel will now have to streamline their activities and do more activities from an online platform. Will business travel and its inherent costs and perceived health risks make sense in the near future? It will also take some time for the markedly contracted aviation industry to expand its operations. New regulations are expected, and a diminished economy will be an obstacle for the airlines to return to the scale of their prior operations. Will the habits that we are developing in the pandemic be permanent? Hopefully, this cleared skied interlude will be a constant reminder of what can be accomplished with decreased carbon emissions and applied to our efforts going forward.

1.6  Historical Impact of Epidemics on the Environment Throughout history, epidemics and the spread of diseases have had an impact on the environment, in particular on carbon emissions. Ice cores examined from the time of the black plague and during the period of European introduction of smallpox to South America have shown decreased levels of carbon dioxide. This may possibly due to large tracts of previously cultivated land returning to the wild. It is unlikely that this will occur as land use will probably remain the same. A more likely correlation will be to that of the reductions in carbon emissions that occurred with the financial crash of 2008 and 2009. The overall drop in emissions was 1.3% per year. This was primarily due to the reduction in industrial activities which accounts for 18.4% of all emissions. The rebound which occurred in 2010 led to all-time high emissions. However, returning to a normal economy is much less certain now [14]. Most importantly, maybe the behavioral changes would carry over beyond the current coronavirus pandemic. Social science research predicts that interventions that take place during moments of change are more effective. Many fundamental changes in long-standing institutions have occurred during or after great upheavals

1 Introduction

7

in society. The formation of the National Health Service in Great Britain following World War II is a good example. Changes are often planned for years in advance, but often a great stimulus is needed to bring it about. Global populations are traveling less, cutting down on food waste, and in general leading a life that is leaving less of a carbon footprint. Time will tell if these changes persist. It is safe to say that no one would want carbon emissions to be lowered this way. The COVID-19 pandemic has taken a profound toll on our lives, health services, and mental well-being. It has also shown the impact that communities can have when looking out for the greater good. That may be one lesson that could be invaluable in dealing with climate change.

1.7  The Blue Economy The world contains approximately 4.5 million fishing vessels, which often remain at sea for weeks or months at a time. Fish is the world’s most widely traded food commodity and continues to be the source of substantial protein during this pandemic. The vulnerability of these operations has been made evident by the rapid spread of COVID-19 within the cruise industry and naval ships. Yet despite these risks, local ports must remain open and the fishing industry must continue. Many fishermen are adapting to these changes by shortening voyages and altering processing protocols. The pandemic has profoundly changed the industry. The demands for different varieties of fish have changed with an emphasis on frozen stocks which can be preserved for longer periods. Much of the industry relied on an intricate system of international processing, ship-to-ship transferring of catch and exportation. In the United States alone, 90% of its seafood is imported or processed overseas, much of the processing occurring in China. Many ports are being closed to foreign vessels. Many distant-water fishing operations have relied on foreign ports for offloading catch. For this reason, much of the fish processing is being transferred to nearby shores to avoid trade disruptions. Countries vary by their ability to sustain these changes with a limited ability to ramp up operations of local fish processing. Illicit fishing is likely to increase as well as law enforcement of the oceans is challenged. A less secure ocean may threaten the sustainability of fish stocks, which is greatly impacted by the enforcement of international regulations. Naval and coastal defense operations are being drawn to other areas of national security, leaving fisheries unchecked. In the near term, fishing operations are likely to decline due to a combination of supply chain complications, workforce declines, and market closures. The last time this occurred at this scale was during the World War II.  During this period, fish stocks rebounded, especially in combat areas such as the North Sea where regional cod and haddock stocks flourished. This was followed by robust catches that persisted well into the 1950s [15].

8

K. Conrad

1.8  Challenges to Operational Ocean Science Research operations from laboratory work to exploratory cruises have been dramatically curtailed because of the pandemic. Climate change has had a profound impact on the oceans, including ice melts, climatic events currents and marine life. Now more than ever, accurate data is needed. Despite efforts to minimize research disruptions at academic and research institutions, most near-term fieldwork has been postponed, and research operations have been limited. The nature of fieldwork makes it difficult to reschedule. Natural events that scientists want to observe often have specific time frames. Research vessels and field stations are often shared by multiple institutions, requiring scientists to plan years in advance. The emphasis for scientific funding now and in the near future, as it should be, will be for direct human health. Scientists from around the world are scrambling to adjust to a rapidly changing environment. Marine researchers, who traditionally rely on fieldwork, are switching to remote observation and learning to present virtually. The University-National Oceanographic Laboratory System (UNOLS), the organization that coordinates oceanographic research vessels across 59 academic institutions, paused operations on its vessels for 30 days starting on March 13. This was extended until July 2020 [16]. Over the next few months and years, delaying fieldwork will delay publications. This could affect policy decisions that are based upon accurate data at a time where it is needed most. Fieldwork, and scientific research in general, have always demanded a high degree of adaptability. Funding is often challenged, and some of the greatest breakthroughs occur at unexpected times. A history of surprise discoveries and disappointing dead ends keep researchers constantly open to make the most of unexpected circumstances. This pandemic presents the greatest of challenges. The global cost of this pandemic will be great, and research will not be spared, but the ingenuity and dedication of the marine scientists will find a way to move the science forward in these challenging times.

References 1. Ripple, W. J., Wolf, C., Newsome, T. M., Galetti, M., Alamgir, M., Crist, E., Mahmoud, M. I., & Laurance, W. F. (2017). World Scientists’ warning to humanity: A second notice. Bioscience, 67(12), 1026–1028. 2. Hodell, D., Curtis, J., & Brenner, M. (1995). Possible role of climate in the collapse of Classic Maya civilization. Nature, 375, 391–394. https://doi.org/10.1038/375391a0. 3. https://www.sciencemag.org/news/2018/06/mysterious-masses-seaweed-assault-caribbean-­­ islands. Retrieved November 9, 2019. 4. Diaz, P. A., Molinet, C., Seguel, M., Diaz, M., Labra, G., & Figueroa, R. I. (2014). Coupling planktonic and benthic shifts during a bloom of Alexandrium catenella in southern Chile: Implications for bloom dynamics and recurrence. Harmful Algae, 40, 9–22.

1 Introduction

9

5. Heisler, J., Glibert, P. M., Burkholder, J. M., Anderson, D. M., Cochlan, W., Dennison b, W. C., Dortch, Q., Gobler, C. J., Heil, C. A., Humphries, E., Lewitus, A., Magnien, R., Marshallm, H. G., Sellner, K., Stockwell, D. A., Stoecker, D. K., & Suddleson, M. (2008). Eutrophication and harmful algal blooms: A scientific consensus. Harmful Algae, 8(1), 3–13. 6. Munro, M. (1999). The discovery and development of marine compounds with pharmaceutical potential. Journal of Biotechnology, 70(1–3), 15–25. 7. Pearson, A. L., Shortridge, A., Delamater, P. L., Horton, T. H., Dahlin, K., Rzotkiewicz, A., et al. (2019). Effects of freshwater blue spaces may be beneficial for mental health: A first, ecological study in the North American Great Lakes region. PLoS One, 14(8), e0221977. 8. COVID-19 pandemic causes dramatic decrease in CO2 levels. https://www.azocleantech.com/ news.aspx?newsID=27214. Retrieved April 28, 2020. 9. Covid-19’s effect on CO2 levels. https://www.mcoscillator.com/learning_center/weekly_chart/ covid-19s_effect_on_co2_levels/. Retrieved April 28, 2020. 10. Polluted cities see clean air and water amid coronavirus shutdown. https://www.mercurynews. com/2020/04/11/photos-polluted-cities-see-clean-air-and-water-amid-coronavirus-shutdown/ 11. Sources of greenhouse gas emissions. https://www.epa.gov/ghgemissions/ sources-greenhouse-gas-emissions 12. https://www.carbonbrief.org/analysis-coronavirus-set-to-cause-largest-ever-annual-fall-­inco2-­emissions 13. United Nations Consensus Report. https://www.Ipcc.ch. Retrieved April 25, 2020. 14. Decrease in CO2 emissions in 2008. https://time.com/3966553/recession-emissions-decline/ 15. World War II and the “Great Acceleration” of North Atlantic Fisheries. https://www.academia. edu/4055383/World_War_II_and_the_Great_Acceleration_of_North_Atlantic_Fisheries. Retrieved April 20, 2020. 16. https://www.unols.org/sites/default/files/US_ARF_Corona_Virus_30day_Revised_Plan.pdf. Retrieved April 20, 2020.

Chapter 2

Changing Ocean Currents David Trossman and Jaime Palter

Time and tide wait for no man. —Geoffrey Chaucer

2.1  Introduction Directly observing ocean currents is a formidable challenge that was done only in small-scale and short-duration field experiments until the advent of satellite observations. In the mid-1970s, satellites first revealed the sea surface temperature of the global oceans, an initial step toward understanding how global surface ocean currents influence temperature patterns at the sea surface. However, the major leap forward in our ability to understand ocean currents was realized when satellite altimeters, launched in the 1990s, provided the first global view of the ocean’s surface circulation, which can be inferred from variations in the height of the sea surface. Nevertheless, the vast ocean interior remained almost entirely hidden from view until a coalition of nations committed to continuously observe the ice-free ocean temperature, salinity, and mid-depth velocity fields with a fleet of more than 3000 robotic drifters called Argo floats in the 2000s. With this explosion of data, we are now able to monitor global ocean temperature, salinity, and circulation in a way that was almost unthinkable less than 30 years ago. However, given that ocean circulation has always experienced substantial temporal fluctuations, and since the observing platforms were launched during a time of rapid climate change [1], these observations are often too short robustly determine whether a warming climate might have already changed ocean circulation patterns. Yet, given the need to underD. Trossman (*) University of Texas at Austin, Oden Institute for Computational Engineering and Sciences, Austin, TX, USA J. Palter Graduate School of Oceanography, University of Rhode Island, Kingston, RI, USA © Springer Nature Switzerland AG 2021 K. Conrad (ed.), From Hurricanes to Epidemics, Global Perspectives on Health Geography, https://doi.org/10.1007/978-3-030-55012-7_2

11

12

D. Trossman and J. Palter

stand ocean circulation and its changes for questions of societal interest, oceanographers have developed a rich suite of tools for the task. The focus of this chapter is on the influence of changes in ocean circulation on consequences of societal interest. For instance, we will consider how changing ocean currents may influence patterns of sea level rise, rather than simply considering rising sea level from all known causes, which is covered in a separate chapter. Due to the challenges in long-term observing and the need for predictions of future changes, numerical modeling has become an indispensable tool used in ocean circulation studies. Observational data are incomplete in space and time and do not often provide all necessary information to infer causal relationships, whereas a numerical models of the climate system—albeit imperfect representations of reality—can be sampled everywhere, at all times, and can be manipulated to test hypotheses. Over time, our representation of the ocean has consistently improved, owing partly to increasing computing power, which offers the capacity to simulate smaller-scale features, including many of the turbulent eddies that are responsible for mixing the ocean. Simultaneously, as observing platforms have improved and in situ data have become more plentiful, we are able to create better physical representations of turbulent processes that still remain too small to be resolved. In what follows, we will first discuss the influence of changing ocean circulation on spatial patterns of sea level rise (Sect. 2.2). Next, we will consider how shifting ocean circulation can influence our climate and daily weather. Section 2.4 will show how fluctuating currents can lead to hotspots of ocean property change, which can have ripple effects on ocean ecosystems having to deal with multiple stressors. We note that this is not an exhaustive list of circulation impacts, and there are important mechanisms that we have left beyond the scope of this chapter. For instance, shifting ocean circulation can concentrate ocean contaminants, including plastics and mercury, in surprising ways [2, 3]. Also, established relationships between ocean circulation changes and changes in major Atlantic hurricane frequency [4, 5] remain poorly understood. One topic in its infancy is our understanding of the roles of ocean temperature and circulation changes on waves [6, 7]. However, it is known that wave effects ultimately influence the likelihood of coastal flooding events when compounded with sea level changes on yearly and longer timescales [8].

2.2  I nfluence of Changing Ocean Currents on Spatial Patterns of Sea Level Globally averaged, sea level rise is proceeding at a rate of more than 3 mm each year. Sea level rise is primarily caused by the addition of melted land-based ice, which increases the ocean’s mass, and by the warming of the ocean, which expands its volume—called the thermosteric effect. The large-scale warming of the ocean (e.g., [9–11]) and associated thermosteric effect was responsible for over 40% of the global average sea level rise from 1993 to 2018 [12]. As sea level rise contribution

2  Changing Ocean Currents

13

from melting glaciers and ice sheets has accelerated in recent years, the fraction of global average sea level rise due to warming has declined [13]. Of great importance to coastal communities, regional sea level fluctuations can be 100 times greater from 1 year to the next than the annual global average rate of sea level rise (Fig. 2.1). These local sea level fluctuations are strongly influenced by ocean circulation, which can pile up water in one place and depress it in another, and/or concentrate warming in a region of heightened ocean heat uptake, thereby locally enhancing the thermosteric effect. A reduction in the saltiness of seawater via, for example, increased precipitation [14, 15], glacial melt [16], or other factors [17–21], also causes volume expansion—called the halosteric effect. Globally averaged, the halosteric term is much smaller than the thermosteric effect, but it can lead to important local sea level fluctuations [22]. Near-coastal sea level variations are driven mostly by the ocean’s dynamic response to atmospheric variations, as has been demonstrated along many continental shelves near populous coastlines. For instance, on the western European shelf, sea level appears to respond sensitively to fluctuations in winds on interannual to decadal timescales [23]. On the shelf along the eastern United States, the seasonal

Fig. 2.1  At the center is a map of the rate of change in sea surface height between 1993 and 2012 from satellite altimetry. Also shown are relative sea level changes (gray lines) for various tide gauges for the period 1995–2012. For comparison, an estimate of global mean sea level change is shown in each panel (red line). (Reproduced from the Intergovernmental Panel on Climate Change Fifth Assessment Report Chap. 13 on Sea Level Rise [97])

14

D. Trossman and J. Palter

cycle in sea level, which can swing by more than 10 cm, is largely governed by the annual cycle of solar heating. However, anomalies relative to that annual cycle can be due to wind changes far away in the open ocean, which propagate to the shelf in slow ocean waves that travel just a few kilometers per day [24]. When these waves approach the shore, their sea level impact can become amplified due to coastline geometry and sloping bathymetry (e.g., [25, 26]). Along the Pacific coast of North America, significant multidecadal variations in the winds slowed the pace of sea level rise relative to the global average from 1993 to almost 2010 [27]. Meanwhile, sea level in the western Pacific rose at a level three times higher than the global mean over a similar period, owing largely to the winds that altered the ocean circulation [28]. In addition to these relatively rapid responses of sea level to the atmosphere, ocean circulation may also shift over longer periods of time and in response to the melting of ice sheets. Much attention has been given to a large-scale system of currents, called the Atlantic Meridional Overturning Circulation (AMOC), because it plays an important role in the climate of the Northern Hemisphere and also can strongly influence sea level patterns on both sides of the Atlantic. Partly because it is a component of the AMOC, the Gulf Stream transports warm water north in the Atlantic [29]. Some of this water makes it all the way to the Nordic Seas between Greenland and Scandinavia, where the warmth is transferred to the atmosphere, and the ocean water becomes extremely cold and dense (Fig. 2.2a). This dense water then spreads southward at great depth. Numerical models and reconstructions of ancient climate changes suggest that the AMOC can slow when the high-latitude North Atlantic seawater becomes less dense, as it does when freshwater is added from the melting ice sheets under global warming [30–33]. A potential future slowdown of this system of currents has important implications for the expected pattern of sea level rise. Model results have linked a projected AMOC slowdown to accelerated sea level rise along the North American coast, including in New  York and Boston [34, 35]. However, observations have recently indicated that much of the correlation during the observational era between AMOC and sea level along the New England coast is due to a shared response of the AMOC and sea level to wind fluctuations [36]. Our record of direct AMOC observations is still less than 20 years old, which is too short to decisively resolve if a slowing in the current system has already helped shape the rate of sea level rise along the US east coast.

2.3  S  hifting Ocean Currents and Predicted Influence on Future Climate In addition to its potential to influence the rate of coastal sea level all along the North American coast, the AMOC is implicated in a number of climate phenomena, including the average temperature of the Northern Hemisphere [31, 37–39]. A slowdown of this current system can cause a reduced rate of Northern Hemisphere

2  Changing Ocean Currents

15

a xw 26°N

xe

0

Cable

3000

Eastern Array

Western Array

6000

Mid-Atlantic Ridge Moorings

b

c

80°N

80°N

40°N

40°N





40°S

40°S

80°S

d

80°S 60°E 120°E 180°W 120°W 60°W

60°E

120°E 180°W 120°W 60°W

0

2

4

6

Temperature aromaly [°C]

Fig. 2.2  Schematic of the Atlantic Meridional Overturning Circulation (top) and the modeled projection of surface air temperature response to an idealized global warming simulation (bottom). The schematic shows the Gulf Stream and its northeastward extension, the North Atlantic Current (NAC) transporting warm water to the Nordic and Labrador Seas where they cool and are transformed to dense water masses that flow near the seafloor. The Deep Western Boundary Current (DWBC) transports the cold, dense waters equatorward. The moorings used to monitor the AMOC strength are depicted in red. This schematic was reproduced from [33]. Panels b–d are from model simulations in which atmospheric CO2 is doubled at a rate of 1% per year. Panel b shows surface air temperature change in response to the CO2 increase in a normal simulation, in which ocean circulation responds dynamically to the warming. Panel c shows the surface temperature change after a doubling of CO2 in a simulation in which the velocity field is replaced with that from a control simulation (i.e., with no change in atmospheric CO2). Panel c shows the zonal average surface air temperature response in the normal model (black) and the model with fixed ocean circulation (blue). The difference between the maps and zonal averages reveals that the change in the large-scale circulation (including a 25% slowdown of the AMOC) slows the pace of global warming, principally due to cooling at high northern latitudes. (Maps and zonal averages reproduced from [39])

warming and even pockets of regional cooling, despite the rising anthropogenic greenhouse gas concentrations in the atmosphere (Fig. 2.2—lower panels). The reason for this connection is largely because the poleward transport of heat and salty water by the current system causes the planet to reflect less sunlight at high latitudes. For instance, the poleward oceanic heat transport pushes Arctic sea ice further north, reducing the surface area covered by ice. Sea ice is an efficient reflector of sunlight, helping to keep the planet cooler. In one idealized climate simulation in which a doubling of atmospheric CO2 triggers an AMOC slowdown of 25%, enhanced reflection of solar radiation causes a temporary cooling over the high-­ latitude North Atlantic of up to 4 °C and in Northern Europe of a few tenths of a degree Celsius [38, 39], even as the global average temperature rises nearly 1.6 °C (Fig. 2.2—lower panels).

16

D. Trossman and J. Palter

However, the potential influence of AMOC on climate is not restricted to temperatures. AMOC changes have also been implicated in a strengthening, southward shift, and eastward extension in the midlatitude storm track [40]. The increased storminess is very likely the result of a slower rate of oceanic heat transport, which causes a stronger sea surface temperature contrast between high and low latitudes in the North Atlantic. This regionally enhanced gradient in ocean temperature may then energize the jet stream and cause a shift in the storm track [41, 42]. The result is that the midlatitude North Atlantic and surrounding continents could experience more storms if global warming causes a slowdown of the AMOC. A separate hypothesized cause of increased wintertime cold-air outbreaks in the Northern Hemisphere midlatitudes is linked to warming of the Arctic, which has approximately doubled the global average pace of warming. This amplified rate of warming is due principally to the melting of reflective sea ice and the opening up of highly absorbing seawater. With this accelerated warming, the temperature contrast between the equator and poles averaged over the Northern Hemisphere has decreased; this is in contrast to the intensification of the North Atlantic sea surface temperature contrast projected to occur during an AMOC slowdown. The weakened hemispheric temperature gradient can lead to a lazy and meandering jet stream, also known as the tropospheric polar vortex. Large meanders of the jet stream can bring anomalously cold air southward and warm air northward, creating extreme weather events at different longitudes (Fig. 2.3). Such meandering can also create wave disturbances that travel upward to the stratosphere where a second polar vortex swirls high above the jet stream, encircling the North Pole in winter [43–45]. The stratospheric polar vortex can then also become wavy and distorted, even splitting into

Fig. 2.3  A schematic of the stratospheric and tropospheric polar vortex, the latter of which is more commonly known as the jet stream. Meanders of these vortices can produce anomalous weather at midlatitudes. An active area of debate is how these dynamical features will respond to warming, with evidence mounting for a wavier jet stream and altered polar vortices. (Modified from [98])

2  Changing Ocean Currents

17

two [44, 45]. Such stratospheric polar vortex distortions can also be driven by Arctic sea ice loss [46]. The transport of heat by the ocean may complicate the relationship between a warmer Arctic and the stability of the jet stream and behavior of the stratospheric polar vortex [47, 48]. Yet, evidence is mounting that changes in the Arctic may be causing extreme weather events in a Northern Hemisphere linked to a more wavy atmospheric circulation.

2.4  T  he Role of Ocean Circulation in Marine Heat Waves, Acidification, and Deoxygenation Marine ecosystems are facing the triple threat of ocean warming, acidification, and deoxygenation, which have been vividly described as the process of ocean water “warming up, turning sour, and losing breath” [49]. As with rising sea levels, the pace of change in these stressors is not uniform across the globe (Fig. 2.4). Shifting ocean circulation can create hotspots of change that exceeds the pace of global averages. One such hotspot is found along the New England and Nova Scotian continental shelf, including in the Gulf of Maine and Gulf of Saint Lawrence. Here, a shift in the large-scale ocean circulation has caused a multidecadal decline in the fraction of cold, oxygen-rich subpolar water on the shelf and an increase in warm,

Fig. 2.4  Future biogeochemical change in the world’s oceans. Maps a–c show the difference between future (i.e., the average from 2091 to 2100) and contemporary (i.e., the average from years 1996 to 2005) values in a future in which CO2 emissions are not reduced under any policy interventions (called RCP8.5). Plots d–f show the global average change relative to contemporary values under RCP8.5 and a future with moderate emission reductions (called RCP4.5) at the ocean surface and seafloor. (Modified from [99])

18

D. Trossman and J. Palter

o­ xygen-­poor subtropical water [50–52]. In the following exploration of the causes and consequences of warming, acidification, and deoxygenation, we will use the New England and Scotian continental shelf as a case study to illustrate how ocean circulation can play a leading role in setting the pace of these disruptive changes. The lion’s share (>90%) of the extra heat retained by our planet due to anthropogenic greenhouse gas emissions is in the ocean [53, 54]. With its great mass and larger density relative to air, this extra heat raises the temperature of the ocean much more slowly than the atmosphere. Globally averaged, the top 75  m of the ocean warmed 0.1  °C between 1971 and 2010, with a smaller warming of 0.015  °C at 700 m depth [55], just fraction of the approximately 0.5 °C of global average surface air temperature warming over the same time period [1]. Because the ocean circulation redistributes heat [38, 56, 57], shifting currents can strongly alter the rate of warming and contribute to the occurrence of marine heat waves. A marine heat wave was recently defined as an event lasting 5 or more days with ocean temperatures warmer than the 90th percentile based on a 30-year historical baseline period [58]. The incidence of surface marine heat waves was shown to double since 1982, when continuous satellite sea surface temperature records became available [59, 60]. The associated ecosystem disturbances have compromised ecological goods and services, such as fisheries landings and coral reef tourism [61]. For example, the New England and Scotian shelves have consistently warmed faster than the global average over the nearly 40-year satellite record, but had a particular temperature surge in the decade starting in 2004. This rapid period of warming, which was very likely tied to circulation changes [51], contributed to the collapse of the Maine cod fishery [62]. In addition to soaking up extra heat, the ocean has also acted as a major sink for anthropogenic CO2, absorbing nearly 30% of all of society’s emissions to date [63]. By absorbing CO2, the ocean slows the buildup of this greenhouse gas in the atmosphere, reducing the pace of global warming. However, the dissolved CO2 is also slowly making the ocean more acidic, with negative impacts on many forms of marine life [64, 65]. Calcifying organisms from the tiny sea butterfly to the reef-­ building corals can be impacted by ocean acidification. Ocean circulation near the coasts has been linked with variations in acidity [66] and the threshold beyond which carbonate-based minerals dissolve [67]. Both the acidity and rates of carbonate-­based mineral dissolution near the bottom of the ocean may be more sensitive to the ocean circulation than almost any other factor [68]. With coral reefs compromised by acidification and warming, their role in dissipating powerful waves will be diminished [69], so low-latitude coastlines may then experience a stronger wave climate in the future. Returning to the New England and Scotian shelves, the pH trend has slightly outpaced the global average but is highly uncertain due to strong variability and a sparse observational record [70]. Concern that this region and its valuable fisheries will be highly sensitive to future change stems largely from the fact that the large inputs of freshwater along the coastline already create a reduced buffering capacity relative to other locales [71]. Ocean deoxygenation is an added threat to marine animals. Warmer water has a lower solubility for gases, including O2. Therefore, just as a carbonated beverage

2  Changing Ocean Currents

19

goes flat as its temperature rises, ocean warming leads to a shrinking reservoir of oxygen for marine animals to breathe [72–74]. This effect explains about half of global ocean deoxygenation [74, 75]. Therefore, other factors also determine where and how fast oxygen is lost from the ocean, many of which can be linked to shifting ocean circulation. The factors contributing to the loss of oxygen from the ocean can be summarized as follows: • Photosynthesis and respiration: Microscopic marine algae, called phytoplankton, produce oxygen through photosynthesis in the sunlit surface ocean. Much of this oxygen is lost from the ocean as the surface seawater equilibrates with the atmosphere. When these algae die, they can be eaten by other organisms or decomposed by microbes. Both of these processes consume oxygen, often at depths in the ocean that do not readily exchange oxygen with the atmosphere. Therefore, changes in the rate of photosynthesis, respiration, and their distribution with respect to depth in the ocean can alter marine oxygen concentrations. With the physical transport of nutrients driving booms and busts in photosynthesis and respiration, shifting ocean circulation may play a critical role in determining the trajectory of this factor. Nutrient cycling has been shown to be locally important in varying ocean oxygen concentrations (e.g., [76]) but is thought to be a small contributor to deoxygenation globally. • Ocean circulation changes: As the large-scale sinking and rising of waters changes, so will the rate at which oxygen can be moved from the surface to deeper layers in the ocean interior. Again, the New England and Scotian shelves provide a striking example of the role of shifting currents in amplifying the stressor of deoxygenation. Here, the shift toward a higher proportion of subtropical waters has been accompanied by a rapid loss of oxygen [50, 52]. This loss has plunged oxygen concentrations to levels known to be stressful to benthic marine organisms. • Mixing changes: Vertical mixing in the ocean brings oxygenation from the surface ocean, where it is near equilibrium with the atmosphere, to greater depths, a process often called ventilation. The rate of vertical mixing is thought to decrease with increased ocean stratification (e.g., [77]). Under anthropogenic global warming, the surface ocean generally warms much faster than deeper layers of the ocean, thus increasing this stratification. Therefore, warming is expected to decrease the downward supply of oxygen to the ocean interior. The rate at which this will effect ocean oxygen concentrations has been inferred to be one of the most important unknowns about how the ocean will deoxygenate [75, 78], and observations have proven that temperature stratification does not always translate to the expected decrease in vertical mixing [79]. Each of these factors is included in our global climate models, with the goal of predicting future levels of oxygen in the ocean. However, ocean models forced with historical data have generally simulated a pace of deoxygenation about half as fast as it has observations suggest it has occurred [74, 80]. This decline is happening fastest in the tropical oceans [81] and some coastal regions [52]. Because of the changes occurring in the tropical oceans and the potential importance of the ocean

20

D. Trossman and J. Palter

circulation changes there, it is essential to be able to capture the currents in those regions to make credible future predictions [82, 83]. Further, to make accurate predictions, models must effectively simulate future changes to winds, which influence how nutrients are transported upward toward the ocean’s surface where they fuel photosynthesis [84–86]. Stronger and/or more frequent tropical storms, for instance, are expected to affect the magnitude and variations of both oxygen concentrations [87] and acidity [88] near the coasts.

2.5  Conclusions There are many ways in which changes in the ocean circulation can have consequences for human health. Sea level is one public health threat due to its ability to displace people [89, 90], and the pace of change experienced on our coastlines can be strongly influenced by shifting ocean currents, in addition to a multitude of other factors. These factors include land ice melt, changes in ocean temperature and salinity, the dynamical adjustment of the oceans to fluctuations in the atmosphere, wave-­ induced effects, imbalances between precipitation and evaporation, and the uplift or sinking of land. We highlighted here the potential role for ocean circulation to modify the sea level that people experience, which can influence the chances of coastal areas being flooded. We also discussed the role of ocean circulation changes on weather and climate, including extreme weather events with potential public health implications. For example, the projected AMOC slowdown may intensify the midlatitude storm track and may play a role in causing extreme temperature and precipitation conditions, especially in Northern Europe. There are also established connections between extreme wintertime temperatures in the Northern Hemisphere midlatitudes and Arctic warming, sea ice decline, and the reduced pole-equator temperature gradient. This connection is thought to be mediated by a destabilized jet stream and ­stratospheric polar vortex. This is disconcerting because such destabilization can cause frigid air outbreaks in the densely populated Northern Hemisphere midlatitudes. Finally, we considered how a shifting ocean circulation may ramp up multiple stressors for marine ecosystems and, by extension, fisheries. Marine heat waves have caused turmoil in many ecosystems (e.g., [91–93]) and even the collapse of fisheries [62, 94]. The increasing acidity of the ocean can influence a variety of marine plants and animals at various life stages. Coral reefs are thought to be especially vulnerable to the combined stressors of warming and acidification. Reefs not only play an important role in ecosystems and tourism but also in buffering the impacts of waves impinging upon coastlines [69]. Finally, the loss of ocean oxygen is driving many species in the ocean to lose habitat [95] and migrate [96]. The role of ocean circulation in accelerating or ameliorating these stressors is an active area of inquiry, with the hope that increased understanding will help create more robust predictions for the future, and a heightened awareness of the interconnected health outcomes for the ocean and society.

2  Changing Ocean Currents

21

References 1. Stocker, T.  F., et  al. (2013). Technical summary. In T.  F. Stocker, D.  Qin, G.-K.  Plattner, M.  Tignor, S.  K. Allen, J.  Boschung, A.  Nauels, Y.  Xia, V.  Bex, & P.  M. Midgley (Eds.), Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge/New York: Cambridge University Press. 2. van der Mheen, M., Pattiaratchi, C., & van Sebille, E. (2019). Role of Indian Ocean dynamics on accumulation of buoyant debris. Journal of Geophysical Research: Oceans, 124, 2571– 2590. https://doi.org/10.1029/2018JC014806. 3. Strode, S., Jaeglé, L., & Emerson, S. (2010). Vertical transport of anthropogenic mercury in the ocean, Global Biogeochem. Cycle, 24, GB4014. https://doi.org/10.1029/2009GB003728. 4. Zhang, R., & Delworth, T. L. (2006). Impact of Atlantic multidecadal oscillations on India/ Sahel rainfall and Atlantic hurricanes. Geophysical Research Letters, 33, L17712. 5. Yan, X., Zhang, R., & Knutson, T. R. (2017). The role of Atlantic overturning circulation in the recent decline of Atlantic major hurricane frequency. Nature Communications, 8, 1695. 6. Reguero, B. G., Losada, I. J., & Méndez, F. J. (2019). A recent increase in global wave power as a consequence of oceanic warming. Nature Communications, 10, 205. https://doi.org/10.1038/ s41467-018-08066-0. 7. Young, I. R., & Ribal, A. (2019). Multiplatform evaluation of global trends in wind speed and wave height. Science, 364, 548–552. 8. Melet, A., Meyssignac, B., Almar, R., & Le Cozannet, G. (2018). Under-estimated wave contribution to coastal sea-level rise. Nature Climate Change, 8, 234–239. https://doi.org/10.1038/ s41558-018-0088-y. 9. Levitus, S., et  al. (2012). World ocean heat content and thermosteric sea level change (0-2000 m), 1955-2010. Geophysical Research Letters, 39, L10603. https://doi. org/10.1029/2012GL051106. 10. Lyman, J. M., & Johnson, G. C. (2014). Estimating global ocean heat content changes in the upper 1800  m since 1950 and the influence of climatology choice. Journal of Climate, 27, 1945–1957. 11. Ishii, M., et al. (2017). Accuracy of global upper ocean heat content estimation expected from present observational data sets. Scientific Online Letters on the Atmosphere, 13, 163–167. 12. Cazenave, A., WCRP Global Sea Level Budget Group, et  al. (2018). Global sea-level budget 1993-present. Earth System Science Data, 10, 1551–1590. https://doi.org/10.5194/ essd-10-1551-2018. 13. Chambers, D. P., et al. (2017). Evaluation of the global mean sea level budget between 1993 and 2014. Surveys in Geophysics, 38, 309–327. 14. Köhl, A. (2014). Detecting processes contributing to interannual halosteric and thermo steric sea level variability. Journal of Climate, 27, 2417–2426. https://doi.org/10.1175/ JCLI-D-13-00412.1. 15. Llovel, W., & Lee, T. (2015). Importance and origin of halosteric contribution to sea level change in the southeast Indian Ocean during 2005–2013. Geophysical Research Letters, 42, 1148–1157. https://doi.org/10.1002/2014GL062611. 16. Lorbacher, K., Marsland, S. J., Church, J. A., Griffies, S. M., & Stammer, D. (2012). Rapid barotropic sea level rise from ice sheet melting. Journal of Geophysical Research, 117, C06003. https://doi.org/10.1029/2011JC007733. 17. Munk, W. (2003). Ocean freshening, sea level rising. Science, 300, 2041–2043. https://doi. org/10.1126/science.1085534. 18. Suzuki, T., & Ishii, M. (2011). Long-term regional sea level changes due to variations in water mass density during the period 1981-2007. Geophysical Research Letters, 38, L21604. https:// doi.org/10.1029/2011GL049326. 19. Llovel, W., & Terray, L. (2016). Observed southern upper-ocean warming over 2005-2014 and associated mechanisms. Environmental Research Letters, 11, 124023. https://doi. org/10.1088/1748-9326/11/12/124023.

22

D. Trossman and J. Palter

20. Volkov, D. L., Lee, S.-K., Landerer, F. W., & Lumpkin, R. (2017). Decade-long deep-ocean warming detected in the subtropical South Pacific. Geophysical Research Letters, 44, 927– 936. https://doi.org/10.1002/2016GL071661. 21. Piecuch, C. G., Bittermann, K., Kemp, A. C., Ponte, R. M., Little, C. M., Engelhart, S. E., & Lentz, S. J. (2018). River-discharge effects own United State Atlantic and Gulf coast sea-level changes. PNAS, 115(30), 7729–7734. https://doi.org/10.1073/pnas.1805428115. 22. Griffies, S. M., & Greatbatch, R. J. (2012). Physical processes that impact the evolution of global mean sea level in ocean climate models. Ocean Modelling, 51, 37–72. 23. Chafik, L., Nilsen, J.  E. Ø., Dangendorf, S., Reverdin, G., & Frederikse, T. (2019). North Atlantic Ocean circulation and decadal sea level change during the altimetry era. Scientific Reports, 9, 1041. https://doi.org/10.1038/s41598-018-37603-6. 24. Calafat, F. M., Wahl, T., Lindsten, F., Williams, J., & Frajka-Williams, E. (2018). Coherent modulation of the sea-level annual cycle in the United States by Atlantic Rossby waves. Nature Communications, 9, 2571. https://doi.org/10.1038/s41467-018-04898-y. 25. Minobe, S., Terada, M., Qiu, B., & Schneider, N. (2017). Western boundary sea level: A theory, rule of thumb, and application to climate models. Journal of Physical Oceanography, 47, 957–977. 26. Wise, A., Hughes, C.  W., & Polton, J.  A. (2019). Bathymetric influence on the coastal sea level response to ocean gyres at western boundaries. Journal of Physical Oceanography, 48, 2949–2964. 27. Bromirski, P. D., Miller, A. J., Flick, R. E., & Auad, G. (2011). Dynamical suppression of sea level rise along the Pacific coast of North America: Indications for imminent acceleration. Journal of Geophysical Research, 116, C07005. https://doi.org/10.1029/2010JC006759. 28. Merrifield, M.  A., & Maltrud, M.  E. (2011). Regional sea level trends due to a Pacific trade wind intensification. Geophysical Research Letters, 38, L21605. https://doi. org/10.1029/2011GL049576. 29. Lozier, M. S., Li, F., Bacon, S., Bahr, F., Bower, A. S., Cunningham, S. A., de Jong, M. F., de Steur, L., de Young, B., Fischer, J., Gary, S. F., Greenan, B. J. W., Holliday, N. P., Houk, A., Houpert, L., Inall, M. E., Johns, W. E., Johnson, H. L., Johnson, C., Karstensen, J., Koman, G., Le Bras, I. A., Lin, X., Mackay, N., Marshall, D. P., Mercier, H., Oltmanns, M., Pickart, R. S., Ramsey, A. L., Rayner, D., Straneo, F., Thierry, V., Torres, D. J., Williams, R. G., Wilson, C., Yang, J., Yashayaev, I., & Zhao, J. (2019). A sea change in our view of overturning in the subpolar North Atlantic. Science, 363, 516–521. 30. Barber, D. C., Dyke, A., Hillaire-Marcel, C., & Jennings, A. E. (1999). Forcing of the cold event of 8,200 years ago by catastrophic drainage of Laurentide lakes. Nature, 400, 344–348. https://doi.org/10.1038/22504. 31. Stouffer, R. J., Yin, J., Gregory, J. M., Dixon, K. W., Spelman, M. J., et al. (2006). Investigating the cause of the response of the thermohaline circulation to past and future climate changes. Journal of Climate, 19, 1365–1387. 32. Palter, J. B. (2015). The role of the Gulf Stream in Europe’s climate. Annual Review of Marine Science, 7, 113–137. https://doi.org/10.1146/annurev-marine-010814-015656. 33. Buckley, M. W., & Marshall, J. (2016). Observations, inferences, and mechanisms of Atlantic meridional overturning circulation variability: A review. Reviews of Geophysics, 54, 5–63. https://doi.org/10.1002/2015RG000493. 34. Yin, J., Schlesinger, M.  E., & Stouffer, R.  J. (2009). Model projections of rapid sea-level rise on the northeast coast of the United States. Nature Geoscience, 2, 262–266. https://doi. org/10.1038/ngeo462. 35. Palter, J. B., Frolicher, T., Paynter, D., & John, J. (2018). Climate, ocean circulation, and sea level changes under stabilization and overshoot pathways to 1.5K. Earth System Dynamics, 9, 817. https://doi.org/10.5194/esd-9-817-2018. 36. Piecuch, C. G., Dangendorf, S., Gawarkiewicz, G. G., Little, C. M., Ponte, R. M., & Yang, J. (2019). How is New England coastal sea level related to the Atlantic meridional overturning circulation at 26°N? Geophysical Research Letters, 46, 5351–5360. https://doi. org/10.1029/2019GL083073.

2  Changing Ocean Currents

23

37. Herweijer, C., Seager, R., Winton, M., & Clement, A. (2005). Why ocean heat transport warms the global mean climate. Tellus, 57, 662–675. 38. Winton, M., Griffies, S.  M., Samuels, B.  L., Sarmiento, J.  L., & Frölicher, T.  L. (2013). Connecting changing ocean circulation with changing climate. Journal of Climate, 26(7), 2268–2278. 39. Trossman, D. S., Palter, J. B., Merlis, T. M., Huang, Y., & Xia, Y. (2016). Large-scale ocean circulation-­ cloud interactions reduce the pace of transient climate change. Geophysical Research Letters, 43, 3935–3943. https://doi.org/10.1002/2016GL067931. 40. Woollings, T., Gregory, J. M., Pinto, J. G., Reyers, M., & Brayshaw, D. J. (2012). Response of the North Atlantic storm track to climate change shaped by ocean-atmosphere coupling. Nature Geoscience, 5, 313–317. https://doi.org/10.1038/ngeo1438. 41. Zhang, R., & Delworth, T.  L. (2007). Impact of the Atlantic multidecadal oscillation on North Pacific climate variability. Geophysical Research Letters, 34, L23708. https://doi. org/10.1029/2007GL031601. 42. Yamamoto, A., & Palter, J. B. (2016). The absence of an Atlantic imprint on the multidecadal variability of wintertime European temperature. Nature Communication, 7, 1093. https://doi. org/10.1038/ncomms10930. 43. Thompson, D.  W., & Wallace, J.  M. (2001). Regional climate impacts of the Northern Hemisphere annular mode. Science, 293, 85–89. https://doi.org/10.1126/science.1058958. 44. Francis, J.  A., & Skific, N. (2015). Evidence linking rapid Arctic warming to mid-latitude weather patterns. Philosophical Transactions of the Royal Society A, 373, 20140170. https:// doi.org/10.1098/rsta.2014.0170. 45. Francis, J. A., & Vavrus, S. J. (2015). Evidence for a wavier jet stream in response to rapid Arctic warming. Environmental Research Letters, 10, 014005. 46. Kim, B.-M., Son, S.-W., Min, S.-K., Jeong, J.-H., Kim, S.-J., Zhang, X., Shim, T., & Yoon, J.-H. (2014). Weakening of the stratospheric polar vortex by Arctic sea ice loss. Nature Communications, 4, 4646. 47. Hu, D., Guan, Z., Tian, W., & Ren, R. (2018). Recent strengthening of the stratospheric Arctic vortex response to warming in the central North Pacific. Nature Communications, 9, 1697. 48. Gastineau, G., L’Hévéder, B., Codron, F., & Frankignoul, C. (2016). Mechanisms determining the winter atmospheric response to the Atlantic overturning circulation. Journal of Climate, 29, 3767–3785. 49. Gruber, N. (1943). 2011: Warming up, turning sour, losing breath: Ocean biogeochemistry under global change. Philosophical Transactions of The Royal Society A, 369, 1980–1996. https://doi.org/10.1098/rsta.2011.0003. 50. Gilbert, D., Sundby, B., Gobeil, C., Mucci, A., & Tremblay, G.-H. (2005). A seventy-two-year record of diminishing deep-water oxygen in the St. Lawrence estuary: the northwest Atlantic connection. Limnology and Oceanography, 50, 1654–1666. 51. Brickman, D., Hebert, D., & Wang, Z. (2018). Mechanism for the recent ocean warming events on the Scotian Shelf of eastern Canada. Continental Shelf Research, 156, 11–22. https://doi. org/10.1016/j.csr.2018.01.001. 52. Claret, M., Galbraith, E. D., Palter, J. B., Bianchi, D., Fennel, K., Gilbert, D., & Dunne, J. P. (2018). Rapid coastal deoxygenation due to large scale ocean circulation shift in the Northwest Atlantic. Nature Climate Change, 8, 868–872. 53. Levitus, S., Antonov, J. I., Wang, J., Delworth, T. L., Dixon, K. W., & Broccoli, A. J. (2001). Anthropogenic warming of Earth’s climate system. Science, 292(5515), 267–270. https://doi. org/10.1126/science.1058154. 54. von Schuckmann, K., Palmer, M.  D., Trenberth, K.  E., Cazenave, A., Chambers, D., Champollion, N., et  al. (2016). Earth’s energy imbalance: An imperative for monitoring. Nature Climate Change, 6, 138–144. https://doi.org/10.1038/NCLIM-15030445C. 55. Rhein, M., Rintoul, S.  R., Aoki, S., Campos, E., Chambers, D., Feely, R.  A., Gulev, S., Johnson, G. C., Josey, S. A., Kostianoy, A., Mauritzen, C., Roemmich, D., Talley, L. D., & Wang, F. (2013). Observations: Ocean. In T.  F. Stocker, D.  Qin, G.-K.  Plattner, M.  Tignor,

24

D. Trossman and J. Palter

S. K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex, & P. M. Midgley (Eds.), Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge/New York: Cambridge University Press. 56. Marshall, J., Scott, J. R., Armour, K. C., Campin, J.-M., Kelley, M., & Romanou, A. (2015). The ocean’s role in the transient response of climate to abrupt greenhouse gas forcing. Climate Dynamics, 44(7–8), 2287–2299. 57. Garuba, O. A., Lu, J., Liu, F., & Singh, H. A. (2018). The active role of the ocean in the temporal evolution of climate sensitivity. Geophysical Research Letters, 45(1), 306–315. 58. Hobday, A. J., Alexander, L. V., Perkins, S. E., Smale, D. A., Straub, S. C., Oliver, E. C. J., Benthuysen, J.  A., Burrows, M.  T., Donat, M.  G., Feng, M., Holbrook, N.  J., Moore, P.  J., Scannell, H.  A., Gupta, A.  S., & Wernberg, T. (2016). A hierarchical approach to defining marine heatwaves. Progress in Oceanography, 141, 227–238. https://doi.org/10.1016/j. pocean.2015.12.014. 59. Frölicher, T. L., Fischer, E. M., & Gruber, N. (2018). Marine heatwaves under global warming. Nature, 560, 360–364. 60. Oliver, E. C. J., et al. (2018). Longer and more frequent marine heat waves over the past century. Nature Communication, 9, 1324. 61. Smale, D. A., Wernberg, T., Oliver, E. C. J., Thomsen, M., Harvey, B. P., Straub, S. C., Burrows, M. T., Alexander, L. V., Benthuysen, J. A., Donat, M. G., Feng, M., Hobday, A. J., Holbrook, N. J., Perkins-Kirkpatrick, S. E., Scannell, H. A., & Moore, P. J. (2019). Marine heatwaves threaten global biodiversity and the provision of ecosystems services. Nature Climate Change, 9, 306–312. https://doi.org/10.1038/s41558-019-0412-1. 62. Pershing, A. J., et al. (2015). Slow adaptation in the face of rapid warming leads to collapse of the Gulf of Maine cod fishery. Science, 350, 809–812. 63. Le Quéré, C., Andrew, R.  M., Friedlingstein, P., Sitch, S., Pongratz, J., Manning, A.  C., Korsbakken, J.  I., Peters, G.  P., Canadell, J.  G., Jackson, R.  B., Boden, T.  A., Tans, P.  P., Andrews, O. D., Arora, V. K., Bakker, D. C. E., Barbero, L., Becker, M., Betts, R. A., Bopp, L., Chevallier, F., Chini, L. P., Ciais, P., Cosca, C. E., Cross, J., Currie, K., Gasser, T., Harris, I., Hauck, J., Haverd, V., Houghton, R. A., Hunt, C. W., Hurtt, G., Ilyina, T., Jain, A. K., Kato, E., Kautz, M., Keeling, R. F., Klein Goldewijk, K., Körtzinger, A., Landschützer, P., Lefèvre, N., Lenton, A., Lienert, S., Lima, I., Lombardozzi, D., Metzl, N., Millero, F., Monteiro, P. M. S., Munro, D. R., Nabel, J. E. M. S., Nakaoka, S.-I., Nojiri, Y., Padin, X. A., Peregon, A., Pfeil, B., Pierrot, D., Poulter, B., Rehder, G., Reimer, J., Rödenbeck, C., Schwinger, J., Séférian, R., Skjelvan, I., Stocker, B. D., Tian, H., Tilbrook, B., Tubiello, F. N., van der Laan-Luijkx, I.  T., van der Werf, G.  R., van Heuven, S., Viovy, N., Vuichard, N., Walker, A.  P., Watson, A. J., Wiltshire, A. J., Zaehle, S., & Zhu, D. (2018). Global carbon budget 2017. Earth System Science Data, 10, 405–448. https://doi.org/10.5194/essd-10-405-2018. 64. Feely, R. A., Sabine, C. L., Lee, K., Berelson, W., Kleypas, J., Fabry, V. J., & Millero, F. J. (2004). Impact of anthropogenic CO2 on the CaCO3 system in the oceans. Science, 305, 362– 366. https://doi.org/10.1126/science.1097329. 65. Feely, R. A., Doney, S. C., & Cooley, S. R. (2009). Ocean acidification: Present conditions and future changes in a high-CO2 world. Oceanography, 22, 36–47. https://doi.org/10.5670/ oceanog.2009.95. 66. Feely, R. A., Sabine, C. L., Hernandez-Ayon, J. M., Ianson, D., & Hales, B. (2008). Evidence for upwelling of corrosive “acidified” water onto the continental shelf. Science, 320, 1490– 1492. https://doi.org/10.1126/science.1155676. 67. Harris, K.  E., DeGrandpre, M.  D., & Hales, B. (2013). Aragonite saturation state dynamics in a coastal upwelling zone. Geophysical Research Letters, 40, 2720–2725. https://doi. org/10.1002/grl.50460. 68. Sulpis, O., Boudreau, B.  P., Mucci, A., Jenkins, C., Trossman, D.  S., Arbic, B.  K., & Key, R.  M. (2018). Current CaCO3 dissolution at the seafloor caused by anthropogenic CO2. Proceedings of the National Academy of Science of the USA, 115(45), 11700–11705. https:// doi.org/10.1073/pnas.1804250115.

2  Changing Ocean Currents

25

69. Ferrario, F., Beck, M. W., Storlazzi, C. D., Micheli, F., Shepard, C. C., & Airoldi, L. (2014). The effectiveness of coral reefs for coastal hazard risk reduction and adaptation. Nature Communications, 5, 3794. https://doi.org/10.1038/ncomms4794. 70. MacLean, M., Breeze, H., Walmsley, J., & Corkum, J. (Eds.) (2013). State of the Scotian Shelf Report. Canadian Technical Report of Fisheries and Aquatic Sciences 3074. 71. Gledhill, D. K., White, M. M., Salisbury, J., Thomas, H., Mlsna, I., Liebman, M., Mook, B., Grear, J., Candelmo, A. C., Chambers, R. C., Gobler, C. J., Hunt, C. W., King, A. L., Price, N. N., Signorini, S. R., Stancioff, E., Stymiest, C., Wahle, R. A., Waller, J. D., Rebuck, N. D., Wang, Z.  A., Capson, T.  L., Morrison, J.  R., Cooley, S.  R., & Doney, S.  C. (2015). Ocean and coastal acidification off New England and Nova Scotia. Oceanography, 28(2), 182–197. https://doi.org/10.5670/oceanog.2015.41. 72. Bopp, L., Le Quéré, C., Heimann, M., Manning, A.  C., & Monfray, P. (2002). Climate-­ induced oceanic oxygen fluxes: Implications for the contemporary carbon budget. Global Biogeochemical Cycles, 16(2), 6–1. https://doi.org/10.1029/2001GB001445. 73. Ito, T., Minobe, S., Long, M. C., & Deutsch, C. (2017). Upper ocean O2 trends: 1958-2015. Geophysical Research Letters, 44, 4214–4223. https://doi.org/10.1002/2017GL073613. 74. Schmidtko, S., Stramma, L., & Visbeck, M. (2017). Decline in global oceanic oxygen content during the past five decades. Nature, 542(7641), 335–339. https://doi.org/10.1038/ nature21399. 75. Oschlies, A., Brandt, P., Stramma, L., & Schmidtko, S. (2018). Drivers and mechanisms of ocean deoxygenation. Nature Geoscience, 11, 467–473. 76. Takano, Y., Ito, T., & Deutsch, C. (2018). Projected centennial oxygen trends and their attribution to distinct ocean climate forcings. Global Biogeochemical Cycles, 32, 1329. https://doi. org/10.1029/2018GB005939. 77. MacKinnon, J., Zhao, Z., Whalen, C.  B., Waterhouse, A.  F., Trossman, D.  S., Sun, O.  M., Laurent, L. C. S., Simmons, H. L., Polzin, K., Pinkel, R., Pickering, A., Norton, N. J., Nash, J. D., Musgrave, R., Merchant, L. M., Melet, A. V., Mater, B., Legg, S., Large, W. G., Kunze, E., Klymak, J.  M., Jochum, M., Jayne, S.  R., Hallberg, R.  W., Griffies, S.  M., Diggs, S., Danabasoglu, G., Chassignet, E. P., Buijsman, M. C., Bryan, F. O., Briegleb, B. P., Barna, A., Arbic, B. K., Ansong, J. K., & Alford, M. H. (2017). Climate process team on internal-wave driven ocean mixing. Bulletin of the American Meteorological Society, 98(11), 2429–2454. 78. Palter, J.  B., & Trossman, D.  S. (2018). The sensitivity of future ocean oxygen concentrations to changes in ocean circulation. Global Biogeochemical Cycles, 32, 738–751. https://doi. org/10.1002/2017GB005777. 79. Somavilla, R., González-Pola, C., & Fernández-Diaz, J. (2017). The warmer the ocean surface, the shallower the mixed layer. How much of this is true? Journal of Geophysical Research Oceans, 122, 7698–7716. https://doi.org/10.1002/2017JC013125. 80. Keeling, R. F., Körtzinger, A., & Gruber, N. (2010). Ocean deoxygenation in a warming world. Annual Review of Marine Science, 2(1), 199–229. 81. Stramma, L.  G., Johnson, C., Spintall, J., & Mohrholz, V. (2008). Expanding oxygen-­ minimum zones in the tropical oceans. Science, 320(5876), 655–658. https://doi.org/10.1126/ science.1153847. 82. Montes, I., Dewitte, B., Gutknecht, E., Paulmier, A., Dadou, I., Oschlies, A., & Garçon, V. (2014). High-resolution modeling of the Eastern tropical Pacific oxygen minimum zone: Sensitivity to the tropical oceanic circulation. Journal of Geophysical Research: Oceans, 119, 5515–5532. https://doi.org/10.1002/2014JC009858. 83. Busecke, J. J. M., Resplandy, L., & Dunne, J. P. (2019). The equatorial undercurrent and the oxygen minimum zone in the Pacific. Geophysical Research Letters, 46, 6716. https://doi. org/10.1029/2019GL082692. 84. Deutsch, C., Berelson, W., Thunell, R., Weber, T., Tems, C., McManus, J., Crusius, J., Ito, T., Baumgartner, T., Ferreira, V., Mey, J., & van Geen, A. (2014). Centennial changes in North Pacific anoxia linked to tropical trade winds. Science, 345, 665–668. https://doi.org/10.1126/ science.1252332.

26

D. Trossman and J. Palter

85. Wang, D., Gouhier, T. C., Menge, B. A., & Ganguly, A. R. (2015). Intensification and spatial homogenization of coastal upwelling under climate change. Nature, 518, 390–394. https://doi. org/10.1038/nature14235. 86. Jacox, M. G., Edwards, C. A., Hazen, E. L., & Bograd, S. J. (2018). Coastal upwelling revisited: Ekman, Bakun, and improved upwelling indices for the U.S.  West Coast. Journal of Geophysical Research: Oceans, 123, 7332–7350. https://doi.org/10.1029/2018JC014187. 87. Rabalais, N.  N., Díaz, R.  J., Levin, L.  A., Turner, R.  E., Gilbert, D., & Zhang, J. (2010). Dynamics and distribution of natural and human-caused hypoxia. Biogeosciences, 7, 585–619. https://doi.org/10.5194/bg-7-585-2010. 88. Manzello, D., Enochs, I., Musielewicz, S., Carlton, R., & Gledhill, D. (2013). Tropical cyclones cause CaCO3 undersaturation of coral reef seawater in a high-CO2 world. Journal of Geophysical Research, Oceans, 118, 5312–5321. https://doi.org/10.1002/jgrc.20378. 89. Hauer, M. E., Evans, J. M., & Mishra, D. R. (2016). Millions projected to be at risk from sea-­ level rise in the continental United States. Nature Climate Change, 6, 691–695. https://doi. org/10.1038/nclimate2961. 90. Hauer, M.  E. (2017). Migration induced by sea-level rise could reshape the US population landscape. Nature Climate Change, 7, 321–325. https://doi.org/10.1038/nclimate3271. 91. Wernberg, T., et  al. (2016). Climate-driven regime shift of a temperate marine ecosystem. Science, 353, 169–172. 92. Oliver, E.  C. J., et  al. (2017). The unprecedented 2015/16 Tasman Sea marine heatwave. Nature Communications, 8, 16101. 93. Hughes, T. P., et al. (2017). Global warming and recurrent mass bleaching of corals. Nature, 543, 373–377. 94. Lam, V. W. Y., Cheung, W. W. L., Reygondeau, G., & Sumaila, U. R. (2017). Projected change in global fisheries revenues under climate change. Scientific Reports, 6, 32607. https://doi. org/10.1038/srep32607. 95. Deutsch, C., Ferrel, A., Seibel, B., Pörtner, H.-O., & Huey, R.  B. (2015). Climate change tightens a metabolic constraint on marine habitats. Science, 348(6239), 1132–1135. https://doi. org/10.1126/science.aaa1605. 96. Pinsky, M. L., Worm, B., Fogarty, M. J., Sarmiento, J. L., & Levin, S. A. (2013). Marine taxa track local climate velocities. Science, 341, 1239–1242. 97. Church, J.  A., Clark, P.  U., Cazenave, A., Gregory, J.  M., Jevrejeva, S., Levermann, A., Merrifield, M. A., Milne, G. A., Nerem, R. S., Nunn, P. D., Payne, A. J., Pfeffer, W. T., Stammer, D., & Unnikrishnan, A. S. (2013). Sea level change. In T. F. Stocker, D. Qin, G.-K. Plattner, M. Tignor, S. K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex, P. M. Midgley, & Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change (Eds.), Climate Change 2013: The Physical Science Basis. Cambridge/New York: Cambridge University Press. 98. Waugh, D. W., Sobel, A. H., & Polvani, L. M. (2017). What is the polar vortex and how does it influence weather? BAMS, 98, 37–44. https://doi.org/10.1175/BAMS-D-15-00212.1. 99. Mora, C., Wei, C.-L., Rollo, A., Amaro, T., Baco, A. R., Billet, D., Bopp, L., Chen, Q., Collier, M., Danovaro, R., Gooday, A. J., Grupe, B. M., Halloran, P. R., Ingels, J., Jones, D. O. B., Levin, L. A., Nakano, H., Norling, K., Ramirez-Llodra, E., Rex, M., Ruhl, H. A., Smith, C. R., Sweetman, A. K., Thurber, A. R., Tjiputra, J. F., Usseglio, P., Watling, L., Wu, T., & Yasuhara, M. (2013). Biotic and human vulnerability to projected changes in ocean biogeochemistry over the 21st century. PLoS Biology, 11(10), e1001682. https://doi.org/10.1371/journsal. pbio.1001682.

Chapter 3

Health Consequences of Marine Oil Spills: Lessons Learned from the Deepwater Horizon Accident Kevin Conrad, Rea Cleland, and Nicholas Reyes

There are certain queer times and occasions in this strange mixed affair we call life when a man takes this whole universe for a vast practical joke, though the wit thereof he but dimly discerns, and more than suspects that the joke is at nobody’s expense but his own. —Herman Melville K. Conrad (*) Ochsner Health, New Orleans, LA, USA e-mail: [email protected] R. Cleland · N. Reyes University of Queensland-Ochsner Clinical School, New Orleans, LA, USA © Springer Nature Switzerland AG 2021 K. Conrad (ed.), From Hurricanes to Epidemics, Global Perspectives on Health Geography, https://doi.org/10.1007/978-3-030-55012-7_3

27

28

K. Conrad et al.

3.1  Offshore Oil Exploration Offshore drilling supplies 30% of global oil production [1]. It has proved to be increasingly difficult compared to land-based operations due to the remote and more challenging marine environment. Aquatic operations have necessitated the need to provide more substantial and technologically advanced production facilities. With expansion into marine territory, oil spills from oil tankers and pipelines along with the need for overland transport bring demanding environmental risks to this industry platform. Offshore drilling began in shallow waters about 200 m deep and remained at this level for several decades. In the 1990s, drilling was expanded deeper as methods were advanced and economic factors encouraged it. Deepwater drilling is typically defined as ranges from 400 to 1500 m deep, while ultra-deepwater drilling currently reaches depths of up to 3000 m. Some companies are exploring depths that exceed 4000 m [2, 3]. Offshore wells, although more expensive to build, tend to have a longer lifespan than onshore wells. With varying degrees, all oceans have been utilized to produce oil with the Arctic being the least explored. In the United States, exploration is legally limited to designated areas primarily off the shore of Texas, Louisiana, Mississippi, Alabama, and California [4]. There is an ongoing debate on whether further areas should be opened for exploration. Several factors, such as tourism, environmental impact, and proximity to seismic events as well as hurricanes, have been taken into account to determine the need for exploratory drills. Traditionally, nations have drilling rights 12 nautical miles from the shore. The Law of the Sea, which the United States has signed but not ratified, establishes that member nations control an exclusive economic zone (EEZ) that extends 200 nautical miles (370 km) from its shore [5]. This piece of legislation is a collection of international laws established by the United Nations, governing the rights, duties, and legal jurisdiction of states in maritime environments. It covers areas such as navigational rights, sea mineral claims, and coastal waters jurisdiction while also holding jurisdiction over water, wind, and petroleum energy [5]. The EEZ confers that a nation has territorial rights to explore and produce oil and gas. In areas of confluence within 200 nautical miles of two or more countries, the separating line is drawn equidistant from the shores of the adjacent nations. The exact extent of exclusive economic zones is a common source of conflicts between states over marine waters.

3.2  Source of Marine Oil Petroleum products are deposited in the seas through a variety of both human-made and natural routes. Oil spills occur due to release of crude oil from tankers, offshore platforms, drilling rigs, and wells. Refined petroleum, which can pose a different environmental impact, may be released from oil tankers or shoreline activities. Oil transportation and drilling accidents are two of the many sources of marine oil spills. According to data from the US Coast Guard from 1991 to 2004, 35.7% of

3  Health Consequences of Marine Oil Spills: Lessons Learned from the Deepwater…

29

the oil spilled in the United States came from tank vessels (ships/barges), 27.6% from oil production facilities, 19.9% from non-tank vessels, 9.3% from pipelines, and 7.4% unidentified [6, 7]. Only 5% of the actual oil spills came from oil tankers, while 51.8% came from other kinds of vessels. All areas experience oil spills with the Gulf of Mexico currently having the highest amount of reported incidences [6, 7]. In addition to production-related events, petroleum seeps occur in a place where natural liquid or gaseous hydrocarbons escape to the earth’s surface. This is generally under low pressure or flow and can occur on land or offshore. Petroleum may escape along geological layers, or through fractures and fissures in the earth. In the Gulf of Mexico alone, there are more than 600 naturally occurring oil seeps that leak approximately 1–5  million barrels of oil per year [8]. Oil spills from catastrophic well drilling accidents as opposed to seeps present a unique environmental danger in which a large quantity of oil is released in a relatively short time period. These pulsatile releases of oil may far exceed that due to seeps, production losses or other means.

3.3  Composition of Marine Oil Marine oil, oil that enters the marine environment through natural or man-made accidents, is composed primarily of crude oil and refined petroleum products. Crude oil is formed over thousands of years when large quantities of dead organisms, mostly zooplankton and algae, are buried underneath sedimentary rock and subjected to both intense heat and pressure [9]. Oil is a finite resource with economically feasible extraction far exceeding the rate of natural production. Over time, crude oil forms a complex mixture of many organic compounds that includes straight-chained, branched, cyclic, monocyclic, aromatic, and polycyclic aromatic hydrocarbons. These hydrocarbons are then formed by smaller components of primarily alkanes, cycloalkanes, and various aromatic hydrocarbons. Additionally, other organic compounds contain nitrogen, oxygen, sulfur, and trace amounts of various metals such as iron, nickel, copper, and vanadium. Some oil reservoirs even contain live bacteria. On the other hand, marine oil includes not only petroleum products in the form of crude oil but also various liquid and solid hydrocarbons. Interestingly, the exact molecular composition of crude oil can be traced to a specific location. It varies from formation to formation, but the proportion of chemical elements varies over relatively narrow limits [9, 10].

3.4  Human Toxicity Different oils and petroleum-related products have different levels of toxicity. This is determined by both their in-ground content and their refined properties. The toxicity of crude oil is related to various components and the degree of which these components exist.

30

K. Conrad et al.

Levels of toxicity to humans are influenced by many factors such as route of exposure, duration of exposure, age at exposure, and coexisting medical factors. Petroleum products broken down by its individual components have been shown to contain a variety of agents that have immediate toxic effects on human life as well as long-term and carcinogenic potential. It is thought that significant exposure in both time and body surface area is required to bring about many of these adverse effects [11]. At high rates of exposure, some crude oil components may cause respiratory, hepatic, renal, endocrine, neurologic, hematologic, and other systemic effects. At moderate doses, mutagenic effects may occur. The carcinogens of concern in crude oil are primarily benzene and polycyclic aromatic hydrocarbons (PAHs) [12]. Benzene is a water-soluble chemical compound (1700 mg/l) which may increase the toxic properties of oil. Exposure can be through inhalation or contact. Studies have revealed that benzene exposure is greatest immediately after crude oil rises to the surface in an oil spill. Standards have been established regulating benzene exposure, but each oil spill represents a unique set of circumstance that makes applying those standards problematic. The National Institute for Occupational Safety and Health recommends a limit of up to 100 ppb over an 8-hour workday. These standards apply to trained healthy adults who are often wearing respirators. The risk for non-­trained and the general public may need to be lower. Benzene is a known human carcinogen with a sizeable amount of research documenting its toxicity. It has been linked to various blood dyscrasias, including aplastic anemia, non-Hodgkin’s lymphoma, and leukemia. Evidence suggests that long-term benzene exposure may cause damage to the reproductive organs [13]. PAHs are a category of chemicals that occur in non-processed crude oil and refined petroleum products. They have been linked to a variety of malignancies in the skin, lung, bladder, liver, and stomach, much of which have been discovered in animal model studies. Human health effects from environmental exposure to low levels of PAHs are not well established [14]. All crude oil contains volatile organic compounds (VOCs) that rapidly evaporate into the air, giving crude oil its distinctive odor. Some VOCs are acutely toxic when inhaled and can trigger an immediate respiratory reaction. They are also potentially carcinogenic if exposure is in proximity to a fresh or massive oil spill. This includes early responders working on the spill and those in coastal communities. VOCs pose the greatest threat during the immediate aftermath of an oil spill as their concentration is rapidly reduced in a relatively short period [13–15].

3.5  Epidemiological Studies on Oil Spills Limited epidemiological studies have investigated the long-term impact of exposures from oil spills on the health of the affected population. Most areas of research have focused on the relatively self-limited acute effects of the spill. The health

3  Health Consequences of Marine Oil Spills: Lessons Learned from the Deepwater…

31

effects of eight major oil tanker spills have been evaluated through limited epidemiological follow-up on residents and cleanup workers. Most of these studies provided evidence for an association between exposure to the oil spill and the appearance of acute physical and psychological consequences [16]. Several of these studies had small sample sizes and primarily collected self-reported health information over a limited time frame. It has been difficult to draw any conclusions from these prior studies. Studies examining the long-term health effects have only recently been undertaken. There is growing scientific awareness that oil spills may cause long-term adverse health implications.

3.6  Deepwater Horizon: Lessons Learned The Deepwater Horizon oil spill is the largest spill in marine history. The Deepwater Horizon drilling platform was located 41 miles off the coast of Louisiana, in the Gulf of Mexico. On the evening of April 20, 2010, high-pressure methane rose into the drilling rig, ignited, and exploded, ultimately engulfing the platform. At the time of the explosion, 126 crewmembers were on board. Eleven missing workers stationed on the rig were never found, despite a 3-day US Coast Guard search operation. They are believed to have died in the explosion. Ninety-four crewmembers were rescued by lifeboat, evacuation vessel, or helicopter, and 17 were treated for injuries. After being consumed by flames, the Deepwater Horizon drilling rig sank on the morning of April 22, 2010 [17, 18]. The total estimated volume of leaked oil is approximately 4.9  million barrels (210 million US gal, 780,000 m3). There is a 10% degree of uncertainty in this estimate. Not all of the leaked oil remained in the ocean as a limited amount was recovered [19]. Containing the open well proved to be a difficult task as technology to contain such a massive spill had not yet been developed. Several short-term efforts provided by the ingenuity of the local petroleum industry were undertaken within the first few weeks. Unfortunately, each of these untested methods to contain the well failed. It took nearly 3 months to completely stem the flow of oil. Finally, on September 19, 2010, National Incident Commander Thad Allen declared the flow of oil from the well had been completely stopped and said that it no longer posed a threat to the Gulf of Mexico [19, 20]. Among industry experts, the difficulty of capping the well raised serious concerns about the ability to respond to catastrophic events in a coordinated manner. According to the satellite images at the time, the spill affected 68,000 square miles (180,000  km2) of the ocean. By early June 2010, oil had washed up along 125  miles (201  km) of Louisiana’s coast and expanded to include coastlines of Mississippi, Florida, and Alabama. It was immediately thought that the spill posed a major threat to marine life, birds, and estuaries [19].

32

K. Conrad et al.

A massive, unprecedented response was undertaken following the accident to protect beaches, wetlands, and estuaries from the spreading of oil. These efforts utilized skimmer ships, floating booms, controlled burns, and 1.84 million US gallons (7000 m3) of oil dispersant. As with the capping efforts, much of the technology and certainly the scale of operations was never before encountered by the petroleum industry. By the summer of 2010, approximately 47,000 people and 7000 vessels were involved in response efforts. Many of the workers were from the local areas and had no prior training in oil remediation [21].

3.7  Ecosystem Effects of Deepwater Horizon Oil Spill The Mississippi River Delta houses 40% of US coastal wetlands. This area provides a crucial habitat to coastal species and plays a critical role in maintaining coastal stability. The fragile delta ecosystem is key in preventing erosion and providing economic opportunity to the area. Before the spill, the Mississippi River Delta ecosystem suffered chronic wetland loss and decline [22]. These factors have led to an economic decline in the region. This area certainly was and is susceptible to any further insults. Although the area had experienced localized man-made environmental accidents, the 2010 oil spill was unprecedented in its extensiveness, impacting shorelines from Texas to Florida [19]. The Gulf of Mexico and adjacent wetlands support a rich biodiversity of both national and global importance. This includes seafood stocks, storm protection, and tourism. Significant death rates were documented for species of dolphins, sperm whales, manatees, sea turtles, mangroves, seabirds, oysters, marine vertebrates, sharks, tuna, and shellfish. Species that surface for air such as dolphins, whales, and sea turtles showed elevated rates of onshore stranding [23]. Specific disease rates among bottlenose dolphins were noted to occur with lethal lung conditions and primary bacterial pneumonia, possibly related to their need to breathe air on the surface. Coastal vegetation, a major fabric of seaside ridge integrity, was damaged following the impacts of the oil spill and led to an accelerated rate of coastal erosion. Many of the communities prior to the spill had already suffered significant loss of land to coastal erosion. Further ecological stress is likely to exert pressure on emergent plants for decades to come as the marshes recover. Over concerns for seafood safety, the National Oceanic and Atmospheric Administration began closing fisheries on May 2, 2010. By June 21, closures covered approximately 37% of the Gulf of Mexico (225,290  km2), extending from Atchafalaya Bay, Louisiana, to Panama City, Florida [24, 25]. Louisiana’s commercial fishing industry bore most of the cost of the spill when compared to the four other Gulf states. According to one study, the spill cost the commercial fishing industry somewhere between $94.7 million to $1.6 billion and anywhere from 740 to 9315 jobs in the first 8 months alone [26].

3  Health Consequences of Marine Oil Spills: Lessons Learned from the Deepwater…

33

3.8  Acute Health Impact The Deepwater Horizon oil spill was an unprecedented catastrophe due to the sheer volume of oil spilled and range it dispersed to. It was also uniquely different in how oil leaked from the ocean floor rather than surface land [27]. Emissions of hydrocarbons from the oil slick were identified as the largest source of primary air emissions, while reaction by-products in the atmosphere, including ozone and secondary organic aerosols, were also detected. In addition to the oil-specific compounds, the oil dispersants used elicited community concerns because of the initial lack of information on its composition and sheer volume used. Following the spill, there were anecdotal reports of health problems among cleanup workers and coastal residents, which included flu-like symptoms, unexplained rashes, and anxiety [28, 29]. A 2011 study by the Louisiana State University Health Sciences Center examined the acute health impacts of the spill as compared to prior spills. The study found similar exposure-related symptoms. This included a variety of symptoms related to contact or inhalation. The most common reported symptoms were coughing, wheezing, and shortness of breath; watery, burning, or itchy eyes; stuffy, itchy, burning, or runny nose; throat or skin rash; sores or blisters lasting at least 3 days; severe headaches or migraines; nausea; excessive fatigue or tiredness; diarrhea; and sore throat [30]. Due to the magnitude of the spill, cleanup workers were employed from a variety of backgrounds. It was reported that safety procedures were not always strictly adhered to. Several fishermen working under British Petroleum’s Vessels of Opportunity program were hospitalized. Despite this, tests conducted on offshore workers failed to detect any dangerous levels of a wide array of chemicals, gases, particulates, and metals. For all onshore workers, heat stress and trauma were found to be the greatest threats to worker safety and health [31]. In general, acute health symptoms in humans after the spill were self- limited, and the focus has transitioned to the potential long-term consequences.

3.9  Long-Term Health Consequences of Oil Spills Perhaps the greatest danger of an oil spill on human health is the chronic effects, an area that has not adequately been studied. Oil spills pose many speculative dangers, but none have been definitively identified through epidemiological studies. The magnitude, confined locality, and aggressive response to the Deepwater Horizon oil spill created an opportunity to examine the long-term impact of oil spills in oceans on human health. Funding at unprecedented level was also available through governmental and private agencies. With this in mind, the National Institutes of Health (NIH) established the Gulf Long-term Follow-up (GuLF) Study to follow the health of workers and volunteers

34

K. Conrad et al.

who responded to the 2010 Deepwater Horizon oil spill in the Gulf of Mexico. This is the largest study to date in this field [32]. The GuLF study has and will continue to examine how the health of people involved in different aspects of oil spill response and cleanup may be affected. The study is assessing not only physical health but also mental health, community resiliency, and economic loss. The study will stratify outcomes based upon the particular task undertaken during the response to the oil spill. The first follow-up by the NIH was carried out from 2013 to 2016. More than 19,000 study participants were contacted by telephone. Over 3500 individuals completed a comprehensive clinical research study exam. The exams included tests of lung and neurological function, physiological measurements, additional collection of biological samples, and questionnaires on physical and mental health. The results of this study are still considered preliminary, and definitive results will not be available for several years. Initial publications, focused on respiratory diseases, have found only limited declines in lung function of workers as compared to nonworkers in selected groups. These cohorts will continue to be followed. This study has also made an effort to engage local residents to ensure that scientific goals are aligned with that of the community [32].

3.10  M  ental Health Outcomes: Association with Health Disparities Significant psychological sequelae occur after exposure to oil spills and other environmental disasters. This has been well documented in previous studies, but not well quantified. These symptoms are similar to the psychological stress seen with post-traumatic stress disorder. It was anticipated that due to the magnitude and close proximity to shore, the Deepwater Horizon oil spill would have similar effects. A 2011 study, looking at Florida and Alabama communities, concluded that income loss and the sequelae of that after the spill may have a more significant psychological health impact than exposure to oil on nearby shorelines [33]. In Louisiana, exposure to past environmental accidents, health disparities, and lack of social support were found to be key factors in predicting mental health disorders in coastal residents. As is often the case, the most vulnerable populations seem to be most at risk from environmental accidents.

3.11  Risk Paradigm The current estimates of human health impacts associated with the oil spill may underestimate those without direct contact to contaminated areas and may extend to areas further inland. Even indirect exposure to the events of the Deepwater Horizon oil spill has been associated with adverse mental outcomes. Much of this is mental stress is thought to be related to economic losses (Fig. 3.1).

3  Health Consequences of Marine Oil Spills: Lessons Learned from the Deepwater…

35

Fig. 3.1  Risk paradigm

Health Disparities

Disaster

Unique Vulnerability

Environmental Threats

3.12  Impact on Woman and Children A unique and related group at risk for mental and physical disorders following oil spills is women and their children. In 2010, the Women and Their Children’s Health (WaTCH) study was established. This was a 5-year study assessing the short- and long-term physical, mental, and community health effects of the Deepwater Horizon oil spill on women and children. In total, 2600 women and 600 children from Southeastern Louisiana participated in the study. Results revealed that 28% of participants suffered from depression and 23% from excessive fatigue, which correlated with the increase in oil exposure [34].

3.13  Conclusion Marine oil spills will continue to have a significant environmental impact not only on marine life but also on human health. Fortunately, improved safety measures have limited but not eliminated the number of events. Although the petroleum industry is in transition, the potential for catastrophic events such as the Deepwater Horizon oil spill will continue to persist. Studies examining the long-term effects of oil spills on human health are lacking, but for the first time, large-scale epidemiological studies have been undertaken. Hopefully, in the next few years, the results from these studies will lend further insight into the impact oil spills have on human health. Despite the development of alternative fuels, the need for petroleum products will continue, and the debate over which areas should be developed will be contested. As nearshore petroleum reserves are exhausted, environmentally sensitive areas such as the Arctic and deep waters will be considered. These areas pose greater technological challenges that may increase the risk of oil spills. The economic benefits of oil exploration must be cautiously weighed against the unknown impact on the increasingly fragile ocean ecosystems and the unknown consequences on human life.

36

K. Conrad et al.

Resources 1. Mabro, R. (2006). Oil in the 21st century : issues, challenges and opportunities (p.  351). Oxford: Oxford University Press for the Organization of the Petroleum Exporting Countries. 2. Freudenrich, C., & Strickland, J. How oil drilling works. Retrieved 21 September 2019. 3. Ronalds, B.  F. (2005). Applicability ranges for offshore oil and gas production facilities. Marine Structures, 18(3), 251–263. 4. US Minerals Management Service. Federal OCS oil & gas production, PDF file. Retrieved 10 February 2020. 5. Part V – Exclusive Economic Zone, Article 56. Law of the Sea. United Nations. Retrieved 10 November 2019. 6. United States Coast Guard. (2007). Cumulative spill data and graphics. United States Coast Guard. Archived from the originalon 2008-10-08. Retrieved 12 September 2020. 7. International Tanker Owners Pollution Federation. Major oil spills. Archived from the original on 28 September 2007. Retrieved November 2 2019. 8. Scientists find that tons of oil seep into the Gulf of Mexico each year. ScienceDaily. January 27, 2000. Retrieved 7 September 2019. 9. Norman, J.  H. (2001). Nontechnical guide to petroleum geology, exploration, drilling, and production (2nd ed., pp. 1–4). Tulsa: Penn Well Corp. 10. Speight, J. G. (1999). The chemistry and technology of petroleum (3rd ed., rev. and expanded ed.) (pp. 215–216). New York: Marcel Dekker, 543. I. 11. Di Toro, D. M., McGrath, J. A., & Stubblefield, W. A. (2007). Predicting the toxicity of neat and weathered crude oil: Toxic potential and the toxicity of saturated mixtures. Environmental Toxicology and Chemistry, 26(1), 24–36. 12. Montagnolli, R.  N., Lopes, P.  R. M., & Bidoia, E.  D. (2015). Screening the toxicity and biodegradability of petroleum hydrocarbons by a rapid colorimetric method. Archives of Environmental Contamination and Toxicology, 68(2), 342–353. 13. Garte, S., Taioli, E., Popov, T., Bolognesi, C., Farmer, P., & Merlo, F. (2000). Genetic susceptibility to benzene toxicity in humans. Journal of Toxicology and Environmental Health, Part A, 71(22), 1482–1489. 14. Bostrom, C.-E., Gerde, P., Hanberg, A., Jernstrom, B., Johansson, C., Kyrklund, T., Rannug, A., Tornqvist, M., Victorin, K., & Westerholm, R. (2002). Cancer risk assessment, indicators, and guidelines for polycyclic aromatic hydrocarbons in the ambient air. Environmental Health Perspectives., 110(Suppl. 3), 451–488. 15. ‘Volatile organic compounds’ impact on indoor air quality. US EPA. Retrieved 4 April 2019. 16. Middlebrook, A. M., Murphy, D. M., Ahmadov, R., Atlas, E. L., Bahreini, R., Blake, D. R., Brioude, J., de Gouw, J.  A., Fehsenfeld, F.  C., Frost, G.  J., Holloway, J.  S., Lack, D.  A., Langridge, J. M., Lueb, R. A., McKeen, S. A., Meagher, J. F., Meinardi, S., Neuman, J. A., Nowak, J. B., Parrish, D. D., Peischl, J., Perring, A. E., Pollack, I. B., Roberts, J. M., Ryerson, T. B., Schwarz, J. P., Spackman, J. R., Warneke, C., & Ravishankara, A. R. (2011). Air quality implications of the Deepwater Horizon oil spill. Proceedings of the National Academy of Sciences, 109(50), 20280–20285. 17. Deepwater horizon. National Oceanic and Atmospheric Administration. Retrieved 28 June 2019. 18. On scene coordinator report on deepwater horizon oil spill (PDF) (Report). September 2011. Archived (PDF) from the original on 15 September 2012. Retrieved 10 August 2019. 19. Robertson, C., & Krauss, C. (2010, 2 August). Gulf Spill is the largest of its kind, scientists say. The New York Times. Retrieved 10 August 2019. 20. Weber, H. R. (2010, 19 September). Blown-out BP well finally killed at bottom of Gulf. Boston Globe. Associated Press. Retrieved 26 February 2011. 21. Ramseur, J. L., & Hagerty, C. L. (2013, 31 January). Deepwater horizon oil spill: Recent activities and ongoing developments (PDF) (Report). CRS Report for Congress. Congressional Research Service. R42942. Retrieved 13 February 2013.

3  Health Consequences of Marine Oil Spills: Lessons Learned from the Deepwater…

37

22. United States Department of the Interior. (1994). The impact of federal programs on wetlands: A report to Congress. Dept. of the Interior. vol. 2. “Archived copy”. Archived from the original on 30 March 2014. Retrieved 20 May 2019. 23. Conservation letters: Whale and dolphin death toll during deepwater disaster may have been greatly underestimated by Dr. Rob Williams, et al. – Offshore Oil Drilling. Reefrelieffounders. com (30 March 2011). Retrieved 7 April 2019. 24. Summary of information concerning the ecological and economic impacts of the BP deepwater horizon oil spill disaster (PDF). Natural Resources Defense Council. Retrieved 4 November 2016. 25. Norse, E. A., & Amos, J. (2010). Impacts, perception, and policy implications of the BP/deepwater horizon oil and gas disaster (PDF). Environmental Law Reporter, 40(11), 11058–11073. ISSN 0046-2284. Retrieved 22 February 2013. 26. Sumaila, U. R., et al. (2012). Impact of the deepwater horizon well blowout on the economics of U.S. Gulf fisheries. Canadian Journal of Fisheries and Aquatic Sciences., 69(3), 499–510. https://doi.org/10.1139/f2011-171. 27. BP Macondo Well Incident. U.S.  Gulf of Mexico. Pollution Containment and Remediation Efforts (PDF). Lillehammer Energy Claims Conference. BDO Consulting. Archived from the original (PDF) on 21 August 2019. Retrieved 17 February 2013. 28. BP Oil Spill Cleanup Workers getting sick, Exxon Valdez Survivor warns of long-term health effects. Democracy Now!. 7 July 2010. Retrieved 1 June 2013. 29. Schmidt, C. W. (2011). Occupational Health. Study to examine health effects in ‘deepwater horizon’ oil spill cleanup workers. Environmental Health Perspectives, 119(5), A204. 30. Investigation: Two years after the BP Spill, a hidden health crisis festers. The Nation. 18 April 2012. Retrieved 1 June 2012. 31. Health Hazard Evaluation of Deepwater Horizon Response Workers National Institute for Occupation Safety and Health, August 2011. 32. GuLF Study. National Institute of Environmental Health Sciences, 9 September 2012. 33. Investigation: Two years after the BP spill, a hidden health crisis festers. The Nation. 18 April 2012. Retrieved 5 June 2019. 34. Peres, L. C., Trapido, E., Rung, A. L., Harrington, D. J., Oral, E., Fang, Z., Fontham, E., & Peters, E. S. (2016). The deepwater horizon oil spill and physical health among adult women in Southern Louisiana: The Women and Their Children’s Health (WaTCH) study. Environmental Health Perspectives, 124(8), 1208–12-13.

Chapter 4

New England Coastal Cities: The Struggle for a Resilient Future H. Curtis Spalding, Siddhi Nadkarni, Claire Bekker, and Devyn Collado-Nicol

Watching a coast as it slips by the ship is like thinking about an enigma. There it is before you, smiling, frowning, inviting, grand, mean, insipid, or savage, and always mute with an air of whispering, “Come and find out”. —Joseph Conrad

4.1  N  ew England Coastal Cities: The Struggle for a Resilient Future Struggling to recover from long-term economic decline starting in the latter half of the twentieth century, midsized coastal cities in Southern New England (SNE) will face daunting challenges generally associated with climate change. As we enter the third decade of the twenty-first century, the most certain consequences of climate change, namely, sea level rise and extreme rainfall events, will add an immeasurable challenge to restoring the vitality that these cities enjoyed throughout much of the twentieth century. This combination of socioeconomic and socio-ecological stress makes the cities of SNE some of the most vulnerable communities in the Northeast, if not the United States. Therefore, as we consider how urban communities across the United States are developing the capability and capacity to adapt to climate change, it is worthwhile to review how SNE cities are beginning to work on that challenge. H. C. Spalding (*) Institute at Brown University for Environment and Society, Providence, RI, USA e-mail: [email protected] S. Nadkarni Brown University, Health and Human Biology, Providence, RI, USA C. Bekker Brown University, Geology-Biology, Providence, RI, USA D. Collado-Nicol Brown University, Environment and Enterprise, Providence, RI, USA © Springer Nature Switzerland AG 2021 K. Conrad (ed.), From Hurricanes to Epidemics, Global Perspectives on Health Geography, https://doi.org/10.1007/978-3-030-55012-7_4

39

40

H. C. Spalding et al.

Toward that purpose, this chapter will summarize the multi-decade socioeconomic decline that Providence, New Bedford, and Bridgeport have experienced throughout the latter half of the twentieth century. These cities were chosen because, as the three largest cities in SNE, they formerly experienced the highest levels of prosperity in the region but are now facing some of the highest levels of poverty. With each city, we will consider how emerging holistic concepts of resilience are being applied to address this climate adaptation challenge. Located in the three states that define SNE—Rhode Island, Connecticut, and Massachusetts—we will gain insights about emerging state resilience initiatives. The discussion that follows describes the socio-ecological foundation for these resilience initiatives and the prospects for this work going forward.

4.2  Economic Struggles Situated on the Southern New England coastline, the settlement of Providence, New Bedford, and Bridgeport preceded the founding of the United States. They share a history of early growth driven by a colonial economy centered on fishing, agriculture, and trade. Like most Northeastern cities, they experienced a period of rapid growth extending from 1870 to 1920 driven largely by rapid industrialization. Nationally, from 1920 to 1950, the economic growth for first-generation American cities slowed down. Inward migration of more racially diverse groups led to exclusionary practices that denied many opportunities to live in desirable neighborhoods and pursue more financially rewarding employment [1]. By the mid-twentieth century, the fortunes of all three cities started to decline. This decline was evidenced by “descriptive decline” and “functional decline,” as framed by Bradbury in Urban Decline and the Future of American Cities. Descriptive decline is characterized by population loss, while functional decline refers to a socially undesirable change that leads to a reduction in the ability to perform social functions—including “large-­ scale production of goods and services,” creative innovation, provision of desirable residential development, and provisions of a social support system for residents [2]. The differences in how each city responded to growth and decline were significant. Whereas Providence and New Bedford were early settlements in New England, Bridgeport was not actually chartered until the nineteenth century. And while the economies of Providence and Bridgeport shifted entirely toward trade and manufacturing as time progressed, New Bedford retained a strong fishing industry. The manufacturing histories of the cities differed too. As the largest and oldest of the three cities, Providence benefited from Rhode Island’s prosperity after the Revolution and early progression toward industrialization [3]. On its way to becoming the contemporary “Silicon Valley” of its time for manufacturing and the jewelry-­ making capital of the country, the city’s population grew to an all-time high of 253,504  in 1940 [4]. New Bedford saw similar growth in the early days of the twentieth century and grew to 121,217 residents by 1920 [5]. Famous for its

4  New England Coastal Cities: The Struggle for a Resilient Future

41

l­eadership in the textile industry, New Bedford boasted an impressive 70 mills, employing over 41,000 workers [4]. By 1920, New Bedford was renowned as one of the richest cities per capita in the United States, and the city opened the New Bedford Textile School to attract skilled labor, designers, and textile chemists; this school would eventually become what we know today as the University of Massachusetts Dartmouth [6]. Unlike Providence and Bridgeport, New Bedford’s growth stalled, and its population slowly declined through most of the twentieth century. Despite a slower start, Bridgeport grew very quickly after new rail lines connected the city to New York, and when munitions manufacturing associated with both World Wars drove rapid growth. Notably, Remington Arms and General Electric, among other significant twentieth-­century corporations, operated large facilities in Bridgeport. Bridgeport’s population peaked in 1950 at 158,709 [7], and while not declining as precipitously as Providence, “functional decline” in Bridgeport was no less impactful [8]. After World War II and throughout the latter half of the twentieth century, the socioeconomic conditions for these cities drastically declined. Public policy driving suburbanization and the relocation of manufacturing to less expensive and less well-­ organized labor markets are generally regarded as the factors driving this decline. During the closing decades of the twentieth century, access to foreign labor markets put even more pressure on the manufacturing economy of New England (Fig. 4.1). During this period, the overall populations between the three cities plummeted by nearly 25%, with Providence experiencing the largest decline.

Fig. 4.1  Total population decline across Providence, New Bedford, and Bridgeport

42

H. C. Spalding et al.

Since 2000, all three cities have modestly reversed the pattern of “descriptive decline” and, taken together, have managed to increase their populations by more than 3.7%. However, given that current population levels are still significantly below the 1960 levels, measures of serious “functional decline” continue to persist. In 2015, the percentage of Providence families below the poverty line was 24.5% compared to an average of 12.3% for “Working Cities”1 across New England [9]. The percentage of families below the poverty line in New Bedford was 19.7%. For comparable “Gateway” Cities in Massachusetts, the percentage was 14.3%. In New Bedford, poverty rates are highest along the coastline of the city, reaching upward of 40%, with most remaining heavily populated with immigrant communities today (Fig. 4.2). For Bridgeport, the poverty rate exceeded 20%, compared to a “Working City” rate of 12.3%. Like New Bedford, Bridgeport’s highest poverty levels are in vulnerable coastal areas. In the South End of Bridgeport, the poverty rate exceeds 50% and the unemployment rate exceeds 35% [11]. For these cities, rates of educational

Fig. 4.2  Percentage in poverty in Bristol County, MA [10]

1  The term “Working Cities” refers to cities that emerged as manufacturing hubs in the nineteenth and twentieth centuries but presently face struggling economies and poor infrastructure. “Gateway Cities” are cities marked by strong state-centered urban economies that often serve as places of opportunity for newcomers [47].

4  New England Coastal Cities: The Struggle for a Resilient Future

43

Table 4.1  Demographic indicators in Providence, New Bedford, and Bridgeport vs. national average (Census Bureau) Demographic indicators Poverty rate % foreign born Median household income Language other than English spoken at home % high school graduate or higher % bachelor’s degree or higher

Average across three cities 23.60% 26.50% $44,944.00 44.80%

National average 12.30% 13.40% $61,372.00 21.30%

75.80% 21.80%

87.30% 30.90%

attainment and measures of mortality and health also underperform when compared to similar cities [12]. As can be seen in Table 4.1, when compared to most municipalities across New England and the nation as a whole, these cities face substantial socioeconomic and demographic challenges toward building more resilient communities. Through the latter decades of the twentieth century and into the twenty-first century, the federal government and the New England states launched programs to arrest the economic decline in an attempt to foster recovery in these SNEs as well as similarly challenged midsized cities across the United States. These initiatives include the US EPA Brownfields and a variety of programs implemented by the US Department of Housing and Urban Development. The New England states launched programs that include tax incentives for economic development, grants, loans for Brownfield cleanups, and infrastructure investments. Taken together, these programs are a strategic effort to address the “functional decline” that persists in older American cities. Providence, Bridgeport, and New Bedford have aggressively participated in these programs and implemented a series of initiatives meant to address declining populations and increasing socioeconomic stresses. Through a combination of funding from federal, state, and municipal levels, each city has approached its socioeconomic challenges with varying degrees of success. Between 2010 and 2014, as a Federal Entitlement Community, New Bedford received three key grants for its Department of Planning, Housing, and Community Development (DPHCD), meant to help the city reshape its housing and economic situation by allocating funds to economic development, homelessness mitigation, infrastructure improvements, and funding nonprofit organizations in the area doing similar work [13]. Guiding this work is a consolidated plan submitted by the city’s DPHCD and approved by the federal Department of Housing and Urban Development (HUD) [14]. This plan laid out strategies for affordable housing and Brownfield redevelopment programs as well as means of support for small business. Perhaps most notably, the plan also included support for traditional harborside industries while indicating a need to diversify port uses to prevent the region from wholly relying on historically failing industries. Furthermore, Bridgeport has received funding from HUD in response to vulnerable communities affected by Hurricane Sandy. Providence and New Bedford also received Community Development Block Grant

44

H. C. Spalding et al.

(CDBG) program funding. In addition to that, the municipality has received funding from the Home Investment Partnership Program (HOME) designed to increase funds for affordable housing. This has been supplemented by programs such as Housing Opportunities for Persons with AIDS, which aims to address inequality for those suffering from the debilitating disease by allowing participants to apply for financial assistance for rehabilitation and recovery and short-term loans for rent and mortgage payments. All programs are funded through HUD and are set to conclude by the end of 2019 [15]. It is important to note that the full scope of programs and initiatives created by the federal government and states with the purpose to revitalize cities like Bridgeport, Providence, and New Bedford were generally implemented through the established vertical bureaucracies of housing, economic development, and environmental protection. As we describe below, resilience planning requires a more holistic systems thinking approach. This foundational aspect of resilience work offers Providence, New Bedford, and Bridgeport an opportunity to develop localized strategies and actions that may have a greater impact on addressing the full scope of socioeconomic and ecological concerns that compromise resilience.

4.3  Sea Level Rise and Extreme Rain Events At the turn of the century, a second challenge became apparent that will dramatically increase the vulnerability of small and midsized coastal cities in Southern New England. Beset by above worldwide average rates of sea level rise and warming ocean temperatures, these cities face the specter for more damaging coastal storms and chronic challenges associated with worsening tidal flooding and coastal ecosystem change. Two of the cities, namely, Providence and New Bedford, have hurricane barriers that afford protection from storms up to what is projected to occur during a Category 3 hurricane. These barriers were built after Hurricane Carol, the 1954 storm that ravaged Southern New England. They were built to protect the cities from storms that have historically had a one in one-hundred-year chance of occurring in any single year. Recently, both barriers are used to prevent flooding during extreme high tide events. The continuing effectiveness of these barriers is an important consideration for flood protection from storms and eventually tides as sea level rise increases especially after 2050. Figure 4.3 shows the possible global sea level rise scenarios up to 2100 [16]. Estimates of sea level rise vary, but models consistently indicate that sea level rise will accelerate through the end of the twenty-first century. Models also show that coastal New England will be disproportionately impacted by sea level rise and storm surge compared to other regions of the country. A study by the US Geological Survey found that the sea levels on the Atlantic coast of North America rose 3–4 times faster than the global average between 1950–1979 and 1980–2009 [17], and with local land subsidence, this trend will continue (Fig. 4.4). Moreover, storm surges and heavy precipitation present further challenges to these cities. Between 1958 and 2011, the number of days with very heavy

4  New England Coastal Cities: The Struggle for a Resilient Future

45

Fig. 4.3  Global sea level rise scenarios for the United States National Climate Assessment (2018) United States Department of Commerce. https://nca2018.globalchange.gov/chapter/2/. Accessed 27 July 2020

Fig. 4.4  Observed change in very heavy precipitation. (Taken from United States National Climate Assessment (2014) United States Department of Commerce. https://nca2014.globalchange.gov/report/our-changing-climate/heavy-downpours-increasing. Accessed 27 July 2020)

46

H. C. Spalding et al.

p­ recipitation increased by 71% in the Northeast, a much higher rate than other regions of the country [18]. Consequently, the Massachusetts Hazard Mitigation Plan listed flooding as the most significant hazard facing the region [19]. The combined climate threats of sea level rise, hurricane risk, storm surges, and heavy precipitation pose a significant hazard to each coastal city’s infrastructure systems. All three cities have both separate storm sewer and combined storm sewer systems. The design and construction of these systems date back to the late 1800s. Because combined sewer systems collect both storm runoff and sewage in single water and wastewater conveyance system, the cities are required under state and federal enforcement mandates and infrastructure planning requirements to collect and treat combined sewer wastewater. As a result, heavy rain events particularly threaten cities with combined sewer systems because overflows are violations of the federal Clean Water Act. The total cost of combined sewer investment for the three cities exceeds 2.0 billion dollars. The CSO (Combined Sewage Overflow) program in Providence alone is estimated to exceed $1.3 billion [20]. Stormwater discharges are already causing or contributing to at least 55% of water impairments in Massachusetts [22]. New permitting requirements associated with water quality impairments in Massachusetts and Connecticut will require ­significant investment in separate stormwater systems. In Providence, the storm drainage infrastructure was based on rainfall estimates from the 1960s and has not been updated to account for heavier precipitation and increased urbanization—putting the city at greater risk for flooding [21]. Furthermore, a revised urban stormwater permit is due for release in Rhode Island and will likely require large infrastructure investments. These costs will add to the enormous investments already committed to upgrading combined sewers. Unfortunately, the infrastructure challenges extend beyond stormwater. Long-­ term sea level rise projections and more intense hurricane predictions associated with warmer water temperatures will eventually put coastal wastewater treatment plants at great risk. The major wastewater treatment plants in Providence and New Bedford are located outside of the cities’ hurricane barriers and are especially vulnerable. Wastewater treatment plant flooding would inhibit discharge and potentially cause backups of raw sewage into residents’ basements and streets [23]. The hurricane barriers in Providence and New Bedford are not equipped to handle rising sea levels and more severe storms. In the case of a Category 2 storm with 4 feet of sea level rise or a Category 3 hurricane at the current sea level, New Bedford’s storm barrier will be compromised [24]. Flooding could also mobilize and spread the polluted soils and sediments in New Bedford’s harbor, worsening environmental health concerns for its most vulnerable residents. Depending on the hurricane category, a number of critical facilities could be impacted in Providence, especially universities, hospitals, ports, transportation, and wastewater infrastructure. With an uncertain climate future, making the infrastructure investments needed to secure Providence, New Bedford, and Bridgeport will be an enormous challenge. By

4  New England Coastal Cities: The Struggle for a Resilient Future

47

focusing on building resilience, as we will describe in the sections that follow, Bridgeport, Providence, and New Bedford could potentially transform how infrastructure investment more optimally enables long-term climate adaptation, compliance with state and federal environmental requirements and secures a fiscally sound future.

4.4  Developing Strategies for Resilience As noted in the previous sections, the challenges facing SNE cities are long term, extraordinarily difficult, and perhaps unprecedented in the era of modern urban development. In response, these cities are embracing the use of resilience thinking and planning to address the impending challenges. This work ranges from using existing planning tools generally associated with emergency response, environmental protection, and community development programs to creating a new group of initiatives that explicitly frame long-term community building around more holistic approaches that address the chronic challenges to community well-being. In the wake of the recent major storms and a growing understanding of the increasing pace of climate change, there has been a fundamental shift in planning methods and goals. This shift is connected to the application of frameworks of thinking derived from socio-ecological concepts of resilience. A simplified application of socio-ecological concepts has driven the improvement of established planning frameworks and tools. These improvements include but are not limited to visualizations that synthesize information about climate change impacts and a more iterative and expanded approach to community engagement. “Clearinghouses” with case studies describing the experience of other communities are now readily available. Recent examples are Rhode Island’s “Storm Tools,” which predicts how sea levels will rise at neighborhood scale [45], and the EPA ARC-X tool [25]. Other examples include the EPA Region 1 Resilience and Adaptation in New England web portal [26] and the Georgetown Climate Center Adaptation Clearinghouse [27]. All of these information transfer tools generally facilitate the ability to consider how climate change will impact a city at a neighborhood scale and enhance communication with political leadership and community engagement. Going beyond the enhancement of FEMA mitigation planning and other existing planning programs, highly prescriptive holistic models for resilience work have emerged that are explicitly grounded in socio-ecological theory. Walker and Salt describe resilience thinking as an “approach to managing natural resources that embraces human and natural systems as complex systems continually adapting through cycles of change” [28]. Lerch in Community Resilience Reader describes a more complex socio-ecological model of resilience:

48

H. C. Spalding et al. Efforts to build community resilience often focus on growing the capacity to “bounce back” from disruptions, like those caused by climate change. But climate change is not the only crisis we face, nor is preparing for disruption the only way to build resilience. Truly robust community resilience should do more. It should engage and benefit all community members, and it should consider all the challenges the community faces, from rising sea levels to a lack of living wage jobs. In addition, it should be grounded in resilience science, which tells us how complex systems—like human communities—can adapt and persist through changing circumstances [29].

The resilience thinking approach described by Lerch and Walker is clearly evident in the conceptual organizing framework for the 100 Resilient Cities initiative founded by the Rockefeller Foundation [30]. Although in the process of closing down, 100 Resilient Cities substantially accelerated resilience building within the networks of major cities that participated. 100 Resilient Cities defines urban resilience as “the capacity of individuals, communities, institutions, businesses, and systems within a city to survive, adapt, and grow no matter what kind of chronic stress and acute shocks they experience.” 100 Resilient Cities further describes building urban resilience as a holistic enterprise that embraces systems thinking approach: “By strengthening the underlying fabric of a city and better understanding the potential shocks and stresses it may face, a city can improve its development trajectory and the well-being of its citizens” [30]. While resilience is most often framed in the context of stresses and shocks associated with climate change, 100 Resilient Cities explicitly recognizes that there are other potential disrupting short- and long-­ term drivers of change that can greatly damage the well-being of a city. As previously described, the rapid decline of manufacturing in Southern New England cities was a shock that the cities were not prepared for. As we have described above, with respect to almost every measure of social and economic well-being, Providence, New Bedford, and Bridgeport experienced decades of precipitous decline. The 100 Resilient Cities initiative ambitiously asserts that cities can do better than declining for decades by explicitly valuing resilience and building the capability to address the social, economic, and structural conditions that compromise resilience.

4.5  The Boston Example As one of the 100 cities chosen to be part of the 100 Resilient Cities initiative, Boston fully embraced the 100 Resilience framework for building resilience. Through their Resilience and Racial Equity Lens strategy, Boston seeks to better understand and alleviate the effects of racism and inequality on human well-being, access to resources, and the distribution of aid during climate-related disasters, among many other issues. With public engagement and sustained input from those in vulnerable communities, Boston has developed projects designed to foster cohesive communities that will more effectively address social challenges and drivers of inequality. In contrast to traditional frameworks of resilience building, like updating flood barriers, Boston’s resilience plan strives to provide greater economic ­pathways

4  New England Coastal Cities: The Struggle for a Resilient Future

49

to success and higher incomes through increasing access to education and job revitalization. Consistent with the tenets of 100 Resilient Cities, Boston recognizes that an integrated effort which addresses the social, economic, and environmental stresses is not only beneficial but essential to long-term resilience building. Unlike the economies of Providence, New Bedford, and Bridgeport, Boston’s economy is thriving. Participation in 100 Resilient Cities has afforded Boston a capacity level that these SNE cities simply don’t have. However, Boston remains relevant to this discussion as it serves as a foundational model upon which other coastal cities in Southern New England can build their own tailored plans, incorporate tools and techniques to increase social accountability, address vulnerabilities, and foster greater engagement within communities. The following sections review how, without the economic vitality and the support of 100 Resilient Cities, Providence, Bridgeport, and New Bedford are addressing the same climate adaptation challenges facing Boston.

4.6  M  oving Resilience Forward in Southern New England Cities SNE cities have taken a diversity of approaches to address resilience and climate change, but at the baseline of each plan lies Federal Emergency Management Agency (FEMA) Mitigation Planning, a set of resources and tools for preparation and recovery in regard to natural disasters and emergencies. The SNE cities addressed in this chapter all lie within FEMA’s Region I [31]. Recent guidance from FEMA prescribes that communities adopt a “Whole Community” approach for mitigation planning that considers and integrates a wider scope of vulnerability concerns. The “Whole Community” approach requires engagement with multiple offices within local government, local businesses, schools, and nonprofits. This guidance clearly demonstrates a shift toward greater socio-ecological thinking within the structure of the FEMA mitigation planning program. Although recent iterations of FEMA mitigation and “multi-hazard” planning have embraced the “Whole Community” approach, the plans generally remain focused on addressing infrastructure-related actions that will decrease vulnerability and therefore do not address the social challenges that compromise neighborhood and/or community resilience. Nonetheless, it is important to appreciate that the recent FEMA mitigation work in Bridgeport, Providence, and New Bedford is more community based and inclusive. To pursue resilience-building opportunities that are more aligned with community resilience building as described by Lerch and Walker and advanced by 100 Resilient Cities, Providence, Bridgeport, and New Bedford are all engaged in resilience work that goes beyond the scope of FEMA Hazard Mitigation Planning. These measures generally embrace building neighborhood-scale capability and environmental sustainability typified by the emphasis on nature-based solutions. Integrated

50

H. C. Spalding et al.

consideration of issues associated with housing, transportation, environmental protection, employment opportunity, and other factors that contribute to the well-being of a community generally characterizes this approach to building resilience. Use of much more intensive community engagement strategies and practices also characterizes this more holistic community-centered approach to planning. The general planning frameworks for each city described in this chapter are outlined below, starting with a description of the city’s FEMA Mitigation Plan followed by other significant efforts taken by the city. Although the cities below share similar social and economic challenges, their location in three different states needs to be considered with respect to how urban resilience work is supported. While the timing and resource commitments in Massachusetts, Connecticut, and Rhode Island are indeed different, the states share a commitment to applying socio-ecological concepts to resilience planning. Further consideration of the economic, social, and geophysical factors that affect resilience building in our three Southern New England cities is an opportunity for future research. Our purpose in this paper is to consider how socio-ecological approaches to resilience building are emerging in smaller struggling cities that are severely resource constrained and highly vulnerable to emerging climate change conditions.

4.6.1  New Bedford The 2016 FEMA Local Multi-Hazard Mitigation Plan for New Bedford, Massachusetts, is a comprehensive approach that addresses the current and future climate risks facing the city [32]. This plan was first created by the Pre-Disaster Mitigation Planning Committee, which was established by the Mayor of New Bedford in 2004. For the 2016 Multi-Hazard Plan, the Committee was comprised of individuals across the various areas of public safety, health, human services, port authority, public facilities and infrastructure, environmental services, and elected officials. Fulfilling its core purpose, the plan performs a risk assessment for flood-related hazards, wind-related hazards, geologic hazards, coastal erosion, and overall economic impact, noting how existing New Bedford infrastructure and assets can inform future planning. The plan also outlines key objectives, action steps, a timeline, and the status of proposed projects. Above all, integration across departments and collaboration are emphasized as key components to drive forward the goals of resilience. Recently completed, New Bedford’s 2016 Multi-Hazard Plan clearly acknowledges important social factors in the context of the FEMA mitigation planning program, with the inclusion of demographic trends, such as the aging of the New Bedford population. As such, it stands as an example of how socio-ecological thinking is influencing FEMA mitigation planning in the SNE region. However, demonstrating the need for broader more community-based resilience planning, the priority actions in 2016 Multi-Hazard Plan generally does not include steps that would address the long-term challenges that make parts of the New Bedford community especially vulnerable.

4  New England Coastal Cities: The Struggle for a Resilient Future

51

Addressing the limitations of FEMA mitigation planning, in September 2016, Massachusetts Governor Charlie Baker signed Executive Order 569 which launched a state program to advance municipal resilience efforts in response to climate change [33]. The Municipal Vulnerability Preparedness (MVP) program supports Commonwealth communities through planning grants [34]. With the award of a $165,000 grant, New Bedford was included in the first group of cities to participate in the MVP program [35]. Through the prescribed planning process modeled on the Community Resilience Building framework [36], New Bedford’s MVP process recruited the involvement of a diverse group of stakeholders, established attainable goals for reliance, engaged the community research process in regard to the threats of climate change, and hosted two MVP workshops through the Community Resilience Building process. Serving as a proactive and highly participative form of community involvement, these workshops allowed community members to provide input with specific regard to the most socially and ecologically vulnerable parts of New Bedford. The MVP process identifies key environmental issues to address, among which are intense storms, heat waves, air quality, flooding, and sea level rise. In response to these threats, community members identified key assets of New Bedford, such as the harbor and hurricane barrier, as tools that can be included in initiatives to increase resilience and improve public health. Notably, in contrast to the Multi-Hazard Mitigation Plan, the New Bedford plan includes priority actions that address socioeconomic vulnerabilities, such as housing, transportation, and the local economy [37]. It also emphasizes “nature-based solutions” such as green infrastructure to drive their community forward on resilience. With the completion of the MVP plan, New Bedford will be eligible to receive MVP action grants to implement the recommended actions in the MVP plan. This two-step planning/ action grant framework ensures a continued focus and commitment to resilience building, and it sets Massachusetts apart for establishing a structured approach with strong incentives for resilience work.

4.6.2  Bridgeport Although much weaker when it made landfall in the Northeast in 2012, the devastation caused by Hurricane Sandy greatly increased Bridgeport’s commitment to preparedness and resilience work across the Northeast region. Prior to Sandy, Mayor Bill Finch’s administration launched BGreen 2020. Described as a sustainability plan that ambitiously integrated social, environmental, and economic goals, BGreen 2020 established a socio-ecological vision for Bridgeport’s revitalization [38, 39]. Notably, the Bridgeport Regional Business Council was a full partner in building and supporting the plan. After Hurricane Sandy struck Bridgeport in 2012, the socio-ecological framework that drove the development of BGreen 2020 was again applied with a new emphasis on building resilience. Development of the 2014 Great Bridgeport Region Hazard Mitigation Plan included the use of the Community Resilience Building

52

H. C. Spalding et al.

(CRB) framework [36] to engage a broad section of the community. Developed by the Nature Conservancy Office in Connecticut, the CRB uses a workshop format to facilitate discussion and identification of resilience issues that matter at a neighborhood scale. The use of the CRB process helped Bridgeport implement FEMA guidance generally described as the STAPLEE method, which is a feasibility analysis that considers the social, technical, administrative, political, legal, economic, and environmental factors when evaluating mitigation actions [40]. In 2015, Bridgeport ambitiously chose to participate in the Rebuild by Design competition announced by the US Department of Housing and Urban Development (HUD). Rebuild by Design is an organization grounded in a socio-ecological approach and is aimed at building resilience through a process of substantial urban redesign. Because of its experience with integrated cross-disciplinary thinking and its comfort level with broadly inclusive community engagement, Bridgeport was well positioned to successfully compete for Rebuild by Design federal funding. After Bridgeport developed a proposal with an estimated cost exceeding $250 million, the Connecticut Office of Housing on Bridgeport’s behalf successfully secured $65 million in Community Development Block Grant support to fund a resilience build-out plan in the South End of Bridgeport, home to neighborhoods of over 8000 residents with a poverty rate of over 50%, an unemployment rate of over 35%, and a residential vacancy rate of over 50%. The South End is one of the most socially and ecologically vulnerable parts of Bridgeport if not New England [11]. As described in the recently released Environmental Impact Statement: The Proposed Action consists of three projects located within the South End of Bridgeport, Connecticut  – the Rebuild by Design (RBD) Pilot Project at the former Marina Village public housing site, a Flood Risk Reduction Project on the east side of the South End neighborhood, and a Resilience Center  – all of which would combine to provide stormwater management, dry evacuation routes (dry egress), a coastal flood defense system, and resiliency education to the community [41].

The South End Bridgeport plan will utilize both gray and green infrastructure, in the form of sea walls, pumps, and parks, to mediate the effects of sea level rise. In conjunction with a physical berm that can be made higher as sea level rises, a storm pump, and an elevated road, Bridgeport will take advantage of a historically significant Frederick Law Olmsted coastal park to serve as the first line of defense. Moreover, the plan includes strong measures that recognize the social challenges of the South End. A Resiliency Center will be developed that will not just provide services in the event of another storm but will also support community-based resiliency work that includes job training, development of local food and healthcare resources, and an activity center for seniors and children. Rebuild by Design is a US Department of Housing and Urban Development (HUD) initiative [11]. As the HUD program intended, Bridgeport’s plan fully embeds socio-ecological thinking in its development and design. However, it is important to appreciate that this work is built upon the vision expressed in the BGreen plan and devastating impact Hurricane Sandy had on the frontline communities in Bridgeport. Thanks to Rebuild by Design funding support, Bridgeport will likely implement the largest resilience project in New England. And when the proj-

4  New England Coastal Cities: The Struggle for a Resilient Future

53

ect is complete, it will be important to evaluate how well the project achieves the goal of greatly strengthening the level of socio-ecological resilience in Bridgeport’s South End.

4.6.3  Providence As an update to the 2013 Providence Hazard Mitigation Plan, the Providence Emergency Management Agency (PEMA) recently released the draft of the 2019 Hazard Mitigation Plan, in which past goals are evaluated and the stage is set for future projects [4]. Notable in the Providence plan is its emphasis on the importance of planning for the future, citing that for every $1 spent on planning, $4 are saved from future loss. A key component of the 2019 Draft Plan is the detailed risk and vulnerability assessments supported with formulae and calculations. Additionally, the plan highlights the importance of community engagement, citing the success of the city’s first public workshop, which took the form of an Online Community Survey—available in both English and Spanish to increase accessibility. The plan also highlights the influential role of academic institutions, nonprofits, representatives of state agencies, utility providers, and other interested stakeholders in building great capacity for developing resilience. As such, the plan outlines past protection measures, risk and vulnerability assessments, public involvement, and important information regarding the use of shelters during disasters—a point that was not as emphasized in the other cities’ plans. Further reinforcing Providence’s commitment to the equitable dissemination of information, the plan emphasizes the importance of notification and warning systems that are best practice for reaching the whole population. The PEMA notification system utilizes up-to-date technology to ensure that the community is prepared to handle the more extreme twenty-first-­ century climate-related disasters while working to establish greater equity to ensure that all citizens, regardless of background, are fully prepared and aware of all threats and hazards. As the newest mitigation plan with respect to the three cities, the 2019 Draft Plan includes more actions that can be characterized as “program building” versus identifying a list of infrastructure projects. Action #1 “Create a Stormwater Learning Center in Roger Williams Park” is an example [4]. The scope and nature of the list of actions suggest that while bound by the extensive requirements associated with FEMA mitigation planning, PEMA has pushed the boundaries, allowing the 2019 Providence Hazard Mitigation Plan to embrace a community resilience focus. Providence’s commitment to community resilience was prominently demonstrated by the development of the “Common Threads: Resilient PVD Design and Resilience Team (DART) Plan” completed in 2016. Although done in a relatively short time frame by an experienced volunteer team from the American Institute of Architects and the New England Sustainability Directors Network, the “DART” Plan is clearly aligned with community-based resilience work. The figure to the right illustrates the significant vulnerability factors that exacerbate the challenges of

54

H. C. Spalding et al.

climate change, broadening the scope of what should be considered in fostering a truly resilient strategy. As stated by Jacqui Patterson, Director of the Climate Justice Initiative of the NAACP, in the Resilient PVD Plan, “The effects of climate change threaten everyone, but they do not threaten all people equally” [42] (Fig. 4.5). Emerging from the “DART” Plan, the creation of the Providence Racial Environmental Justice Advisory Committee further deepened the city’s commitment to climate justice in its climate mitigation and adaptation work. This commitment was demonstrated by the city of Providence, as its intention to drive climate change work through the lens of climate justice was announced [43]. Having already established aggressive climate mitigation and adaptation goals grounded in socio-ecological thinking, Providence chose to defer participating in the first iteration of Rhode Island’s Municipal Resilience Program (MRP). Modeled on the Massachusetts Program and using the Nature Conservancy’s Community Resilience Building framework, the Rhode Island MRP initiative embraces a community-­based approach to resilience building [44]. More than Bridgeport and New Bedford, Providence has framed resilience in closest accord with a community-­ based model for resilience planning as described by Walker and Lerch and implemented in Boston. Although resilience work in Boston set a strong example with its focus explicit on equity, Providence, by amplifying the voices of “frontline” communities, has arguably gone further to frame resilience through the lens of equity and the long-term factors that have disadvantaged indigenous and minority communities. However, Providence has not as yet built on the work of the DART ­resilience scoping effort and moved forward with a formal broad-based resilience planning effort that is similar in scope and ambition to the work in New Bedford and Bridgeport. On the one hand, the delay is likely purposeful: Providence wanted to first engage and empower people living in less advantaged neighborhoods of the

Fig. 4.5  Major vulnerability factors to climate change [42]

4  New England Coastal Cities: The Struggle for a Resilient Future

55

community. On the other hand, without funding that follows a storm event (as was the case in Bridgeport) or a commitment for significant funding from the state (as was the case in Massachusetts), finding the resources to complete a broadly scoped resilience plan that builds on the 2019 Hazard Mitigation and DART Plan is a significant challenge for Providence.

4.7  R  esilience Building in Southern New England Cities: Summary and Questions With an increasing number of analytical and engagement tools in use, innovative community-based initiatives are at the forefront of the movement to build resilience against climate change. These initiatives generally consist of a much more inclusive and ambitious approach to hazard mitigation planning in accordance with FEMA requirements, and newly developed resilience planning efforts strive to engage more of the community in highly participative processes that embrace a holistic integration of community resources—encouraging a multifaceted response that addresses the various socioeconomic and environmental vulnerabilities that exist today in older American cities. These vulnerabilities are especially significant in Providence, New Bedford, and Bridgeport—three of the largest and most historic coastal cities located in Southern New England, where sea level rise and more extreme rainfall will increasingly compromise the capability of the infrastructure that’s necessary to sustain the well-being of these cities. After reviewing the resilience work grounded in socio-ecological theory in Providence, New Bedford, and Bridgeport, several questions for further study emerge: 1. When compared to resilience building initiatives in much larger cities, will funding limitations in midsized cities compromise the effective identification of necessary actions to substantially contribute to greater long-term resilience? 2. Will the existing siloed planning frameworks, which are often driven by top-­ down federal and state requirements for housing, economic development, public works, and other aspects of municipal government, fully align with the priorities identified in the resilience plans? 3. Using the cost of seawalls as a proxy for estimating needed investment in protecting coastal communities, recent work by Resilience Analytics suggest that coastal protection costs in Rhode Island, Massachusetts, and Connecticut will exceed 20 billion dollars by 2040 [46]. Will the uncertain federal and state prospects for substantial infrastructure investment discourage continued focus on resilience building? There are numerous other opportunities for continued investigation and research. The most compelling, fundamental question is whether resilience building as described in this chapter will indeed have a transformative effect on the future of

56

H. C. Spalding et al.

older American cities in Southern New England and across the United States. More research considering this question is essential to ensuring that cities have the opportunity to improve their livelihoods as they struggle to adapt to climate change and other forces that will inevitably disrupt the well-being of urban communities throughout the twenty-first century.

References 1. Powers, A. (2010). Phoenix cities: The rise and fall of great industrial cities (pp. 293–298). Portland: Policy Press. 2. Bradbury, L., Downs, A., & Small, K. (1982). Urban decline and the future of American Cities. Washington, DC: The Booking Institution. 3. City of Providence. (2019). http://www.providenceri.gov/. Accessed 16 June 2019. 4. Strategy for reducing risks from natural, human-caused, and technologic hazards in providence, Rhode Island: A multi-hazard mitigation plan. (2019). http://www.providenceri.gov/ wp-content/uploads/2019/03/Draft-Providence-Multi-Hazard-Mitigation-Plan-Update-2019. pdf. Accessed 16 June 2019. 5. New Bedford, Massachusetts Population 2019. (2019). World population review. http://worldpopulationreview.com/us-cities/new-bedford-ma-population/. Accessed 16 June 2016. 6. UMD history and the textile industry, 1895-1947. (2019). UMass Dartmouth. https://www.lib. umassd.edu/archives/umassd-history/umd-history-and-textile-industry-1895-1947. Accessed 16 June 2019. 7. Population of Bridgeport, CT. (2019). Population.us. https://population.us/ct/bridgeport/. Accessed 16 June 2019. 8. Bailey, H. (2015). Where ‘brownfield’ is a pretty word for toxic dump. In: The Atlantic. https:// www.theatlantic.com/politics/archive/2015/09/where-brownfield-is-a-pretty-word-for-toxicdump/432843. Accessed 16 June 2019. 9. Brown, J., Ramsawak, R., & Gomes, J. (2016). Data profiles (p. 78). Rhode Island, Providence: Mosakowski Institute for Public Enterprise. 10. SimplyAnalytics. (2018). EASI percent of population in poverty 2018. Accessed 20 February 2019. 11. Claim the Edge, Connect the Center. (2014). Rebuild by design: Resilient Bridgeport. https:// resilientbridgeport.com/pdf/Final_Presentation.pdf. Accessed 16 June 2019. 12. Mid-sized cities: Data profiles. (2019). Mosakowski Institute for Public Enterprise. https:// www.clarku.edu/centers/mosakowski-institute/mid-sized-cities-data-profiles/. Accessed 16 June 2019. 13. Planning, Housing, and Community Development. (2019). New Bedford. https://www. newbedford-ma.gov/community-development/resources/ohcd-documents/. Accessed 16 June 2019. 14. Consolidated Plan (2010-2014) City of New Bedford, Massachusetts. http://newbedford. wpengine.netdna-cdn.com/community-development/wp-content/uploads/sites/34/2015/20102014_Consolidated_Plan.pdf. Accessed 16 June 2019. 15. Elorza, J. (2015–2019). Consolidated & annual action plan. https://www.providenceri.gov/ wp-content/uploads/2017/05/Planning-2015_2019ProvidenceConsolidatedPlan.pdf. Accessed 16 June 2019. 16. Global sea level rise scenarios for the United States National Climate Assessment. (2012). United States Department of Commerce. https://scenarios.globalchange.gov/sites/default/ files/NOAA_SLR_r3_0.pdf. Accessed 16 June 2019.

4  New England Coastal Cities: The Struggle for a Resilient Future

57

17. Sallenger, A.  H., Jr., Doran, K.  S., & Howd, P.  A. (2012). Hotspot of accelerated sea-level rise on the Atlantic coast of North Atlantic. Nature Climate Change, 2, 884–888. https://doi. org/10.1038/nclimate1597. 18. U.S. Environmental Protection Agency. (2014). EPA New England regional climate adaptation plan. https://www3.epa.gov/climatechange/Downloads/Region1-climate-change-adaptationplan.pdf. Accessed 18 November 2018. 19. City of New Bedford. (2016). City of New Bedford local multi-hazard mitigation plan update. http://newbedford.wpengine.netdna-cdn.com/emergency-management/wp-content/uploads/ sites/28/pdfs/New-Bedford_MA-HMP-Final-052516.pdf. Accessed 18 November 2018. 20. U.S.  Environmental Protection Agency. Region 1: EPA New England—What are combined sewer overflows (CSOs)? https://www3.epa.gov/region1/eco/uep/cso.html. Accessed 21 February 2019. 21. City of Providence. (2019). Strategy for reducing risks from natural, human-caused and technological hazards in Providence, Rhode Island: A multi-hazard mitigation plan. http://www. providenceri.gov/wp-content/uploads/2019/03/Draft-Providence-Multi-Hazard-MitigationPlan-Update-2019.pdf. Accessed 5 April 2019. 22. National Municipal Stormwater Alliance. (2018). 2018 State of stormwater report. http:// nationalstormwateralliance.org/wp-content/uploads/2018/07/State-of-Stormwater-Report.pdf. Accessed 26 February 2019. 23. Kessler, R. (2011). Stormwater strategies: Cities prepare aging infrastructure for climate change. Environmental Health Perspectives, 119(12), a514–a519. https://doi.org/10.1289/ ehp.119-a514. 24. SeaPlan. (2014). Climate change vulnerability assessment and adaptation planning study for water quality infrastructure in New Bedford, Fairhaven, and Acushnet. http://climate.buzzardsbay.org/download/seaplan-climate-vulnerability-new-bedford-area.pdf. Accessed 18 November 2018. 25. Climate Change Adaptation Resource Center. (2018). EPA. https://www.epa.gov/arc-x. Accessed 16 June 2019. 26. Climate change: Resilience and adaptation in New England (RAINE). (2019). EPA. https:// www.epa.gov/raine. Accessed 16 June 2019. 27. Georgetown Climate Center. (2019). Georgetown law. https://www.georgetownclimate.org/. Accessed 16 June 2019. 28. Walker, B., & Salt, D. (2006). Resilience thinking: Sustaining ecosystems and people in a changing world. Washington, DC: Island Press. 29. Lerch, D. (2015). Six foundations for building community resilience. https://www.postcarbon. org/publications/six-foundations-for-building-community-resilience-2/. Accessed 16 June 2019. 30. 100 Resilient Cities. (2019). Boston. http://www.100resilientcities.org/strategies/boston/. Accessed 19 May 2019. 31. Federal Emergency Management Agency. (2016). Region I: CT, ME, MA, NH, RI, VT. https:// www.fema.gov/region-i-ct-me-ma-nh-ri-vt. Accessed 19 May 2019. 32. City of New Bedford. (2016). Local multi-hazard mitigation plan update. http://newbedford. wpengine.netdna-cdn.com/emergency-management/wp-content/uploads/sites/28/pdfs/NewBedford_MA-HMP-Final-052516.pdf. Accessed 19 May 2019. 33. Commonwealth of Massachusetts. (2016). No. 569: Establishing an integrated climate change strategy for the Commonwealth. https://www.mass.gov/executive-orders/no-569-establishingan-integrated-climate-change-strategy-for-the-commonwealth. Accessed 19 May 2019. 34. Commonwealth of Massachusetts. (2019). Municipal vulnerability preparedness (MVP) program. https://www.mass.gov/municipal-vulnerability-preparedness-mvp-program. Accessed 19 May 2019. 35. City of New Bedford Municipal Vulnerability Preparedness Program. (2018). Community resilience building workshop summary of findings report. http://newbedford.wpengine.netdna-

58

H. C. Spalding et al.

cdn.com/environmental-stewardship/wp-content/uploads/sites/39/New-Bedford-Final-MVPSummary-of-Findings-Report.pdf. Accessed 19 May 2019. 36. Community Resilience Building. (2019). The nature conservancy. https://www.communityresiliencebuilding.com/. Accessed 19 May 2019. 37. New Bedford’s plan for community climate action + resilience. (2019). New Bedford. https:// www.newbedford-ma.gov/environmental-stewardship/nb-resilient-climate-action-resilienceplan/. Accessed 16 June 2019. 38. Infrastructure, utilities, & waste. (2019). NB resilient. http://www.newbedford-ma.gov/ environmental-stewardship/wp-content/uploads/sites/39/New_Bedford_FACTSHEET_ Infrastructure_FEB_05.pdf. Accessed 16 June 2019. 39. BGreen 2020. (2010). A sustainability plan for Bridgeport, Connecticut. http://library.rpa.org/ pdf/BGreen-2020.pdf. Accessed 19 May 2019. 40. 2014 Natural hazard mitigation plan update. (2014). Greater Bridgeport Regional Council. http://www.ctmetro.org/uploads/PDFs/Projects/Environment%20and%20Sustainability/ Natural-Hazard-Mitigation/Natural-Hazard-Mitigation-Plan-2014_opt.pdf. Accessed 16 June 2019. 41. National Disaster Resilience and Rebuild by Design Projects. (2019). Bridgeport, Connecticut draft environmental impact statement/environmental impact evaluation. 42. Common threads: Resilient PVD. (2016). Design and Resilience Team. Presented at the NSMSN/AIA DART Presentation, Providence, RI, 3 February 2016. 43. Climate Justice Plan PVD. (2019). City of Providence. http://www.providenceri.gov/sustainability/climate-justice-action-plan-providence/. Accessed 16 June 2019. 44. Folke, C. (2016). Resilience (Republished). Ecology and Society, 4(44), pages 1–30 45. RI Shoreline Change Special Area Management Plan. (2019). Storm tools. http://www. beachsamp.org/stormtools/. Accessed 21 June 2019. 46. Climate Costs in 2040. (2019). Pay up climate polluters. http://www.climatecosts2040.org/. Accessed 21 June 2019. 47. Federal Reserve Bank of Boston. (2019). Research & rationale. In: Working Cities Challenge. https://www.bostonfed.org/workingcities/about/research.htm. Accessed 25 June 2019.

Chapter 5

The Blue Mind Kevin Conrad, Rea Cleland, and Nicholas Reyes

We are tied to the ocean. And when we go back to the sea – whether it is to sail or to watch it – we are going back from whence we came. —John Kennedy

K. Conrad () Ochsner Health, New Orleans, LA, USA e-mail: [email protected] R. Cleland · N. Reyes University of Queensland-Ochsner Clinical School, New Orleans, LA, USA © Springer Nature Switzerland AG 2021 K. Conrad (ed.), From Hurricanes to Epidemics, Global Perspectives on Health Geography, https://doi.org/10.1007/978-3-030-55012-7_5

59

60

K. Conrad et al.

5.1  T  he Blue Gym: Threats to the Mental Health Benefits of the Ocean Mankind is naturally drawn to water when they desire peace, pleasure, and relaxation. It has been a source of recreation, an inspiration for the arts, and, at times, a healer of personal struggles. As the narrator, Ishmael, states in Hermann Melville’s Moby Dick, “whenever it is a damp, drizzly November in my soul, I find it high time to get to the sea as soon as I can.” This sentiment has been expressed in literature, art, and cinema throughout history. The seas are our great consolers. Professionals specializing in this area, aquatic therapists, are increasingly recognizing the healing properties of our oceans and look to the water to help treat a variety of conditions. These conditions include PTSD, addiction, anxiety disorders, and autism [1]. As human beings, we love the water because we subconsciously realize we are part of it, and it is a part of us. However, we also know we can never really grasp the vast unknown that lies below the waters. We intuitively know this connection and know that this connection goes beyond the tactile sensations, pleasant sounds, and nostalgic memories of the days spent at the beach. Something deeper is going on. When we hear the ocean, rivers, and streams call to us, we are hardly able to resist its’ calming properties. What doctors, priests, and spiritual healers have known for millennia, scientists are now documenting. Beyond the spiritual explanations, there are several theories that attempt to define the medicinal and beneficial emotional effects of being near or in water. The term “blue mind” was coined by neuroscientists and marine researchers to identify a personality trait that experiences a full awareness of water and its influence on life [2]. The term describes the somewhat meditative state people fall into when near water. These benefits may be in response to the over-connected and over-­ stimulated state that is modern life. Being devoid of cell phones and computers and the constant 24-hour news cycle in itself may be beneficial, but there may be something uniquely beneficial to being near ocean environments.

5.2  The Blue Gym: Research Outcomes Research on the beneficial effects of water goes beyond the historically documented aspects of this phenomenon and provides biological clues as to why this relationship exists. Experiencing the sound and sites of waves can be described as having visual and auditory inputs with regularity, but without monotony. A wave is predictable and unique at the same time. This has been described as a perfect metaphor to describe the capacity the brain has with enhanced problem-solving and creativity. A similar experience occurs with music where there is a predictable melody, but with nuances that gently challenge the brain [2, 3]. The consciousness is relaxed, but subtle auditory changes allow the brain to be open to new thought patterns.

5  The Blue Mind

61

Several studies have shown a direct correlation between proximity to water and improved overall health. The benefits may not only be limited to mental health; functionality and longevity may be improved as well. While it is important to note the correlation between higher health literacy and social currency with the high cost of living often required for coastal living, the research is still worth considering. In a 2016 study in Wellington, New Zealand, a cross-sectional study found that higher levels of blue space visibility in residential spaces were associated with lower psychological distress. Interestingly, green spaces did not have such an effect. Researchers could attempt to see if other cities with similar natural bodies of surrounding water had a similar impact on mental health [4, 5]. Canadian researchers revealed that living near water reduced the risk of premature death by 12–17% among urban residents. This was particularly true for deaths related to stroke or respiratory-related causes [6]. A European study looking at university students, traditionally under some degree of stress, found that they might benefit from exposure to blue spaces. The students attended an urban institution and were subjected to variable exposures to water environments. The benefits achieved were found to be case-independent of exercise levels [7]. Notably, the benefits of blue spaces may extend beyond oceans to include freshwater blue spaces, such as the North American Great Lakes. In fact, in a 2019 study, both distance to the Great Lakes and percentage of inland lakes had a protective effect on mental health. Interestingly, these protective effects on mental health persisted regardless of the size of the body of water. Additionally, a shorter distance to a Great Lake was correlated with lower anxiety and mood disorder hospitalization rates. Surprisingly, this effect was the opposite for those that lived closer to inland lakes—the population experienced higher rates of hospitalization rates. Perhaps, this is due to the expansiveness and overall sense of vastness that comes with the Great Lakes, as well as the feeling of being fully at one with nature through all sensory modalities. Inland lakes tend to be smaller and tend to be more regulated, limiting overall accessibility [8]. In a September 2019 study, it was revealed that general health in English population is higher among those living closer to the coast. Further, the association was strongest among lower-income groups who may have limited opportunities to improve their mental health. The authors suggested that coastal living may be a way to mitigate some of the mental stress of having a lower income [9]. Proximity and access to bodies of water serve as a major obstacle for certain populations, mainly those of a lower socioeconomic background. Physical barriers to ocean access may overlap with safety concerns and fears of verbal and physical abuse. Organizations such as the Mewater Foundation in San Francisco, California, recognize this intersectionality. They work to address issues of trauma and stress in young people from backgrounds of poverty and violence by providing access to nature and the ocean through surfing. There have been several studies that evaluated the impact of surfing on not only young persons with varying mental health issues but also persons of any age with a disability. A study focusing on the benefits of surfing for children with cerebral

62

K. Conrad et al.

palsy determined that the increase in aquatic exercise was positively correlated with improved muscle strength, cardiorespiratory function, and gross motor skills [10]. Another found significant improvements in overall fitness, greater environmental awareness, and more positive attitudes toward education and their peers. Additionally, this study found that surfing helped improve self-esteem and overall well-being in young persons [10]. It can be inferred that exposure to blue spaces indeed has a positive effect on mental health. Still, further research needs to be conducted to see to what extent said positive effect goes and in what populations would be most benefiting from blue space exposure. Benefits of water exposure may not be limited to natural environments. In a 2017 British Journal of Sports Medicine study, a significant reduction in cardiovascular mortality was observed for participation in swimming not seen in other exercises of equal aerobic activity. They found that participation in a swimming exercise program was associated with 28% lower risk of premature death from any cause. The authors also concluded that intense training was not necessarily needed to enjoy these benefits as they were seen in less strenuous swimming activities as well [11]. There is still so much that has yet to be discovered regarding the oceans themselves and the impact they have on our well-being. Nonetheless, these studies demonstrate just a handful of benefits of exposure to aquatic environments on our population. These benefits occur in variety of situations and across diverse demographic populations.

5.3  Threats to the Blue Gym As ocean and waterway pollution from various sources continue to increase, it can be expected that the mental benefits of our “blue gym” will decline. Direct causes such as inability to spend time near the beach or in the waters will have a significant impact. Much of the Mississippi Gulf Coast was closed in the Summer of 2019 due to algal blooms caused by nutrient runoff from the Mississippi River. The runoff was primarily generated from upstream farmlands [12]. The beach has been a historical source of recreation and commerce for this region, and indeed much of the industry around this area is related to water sports. This was not an isolated event as beach closures due to algal blooms are occurring globally at an increasing rate. The mental, physical, and economic burden of these events is largely unknown, but researchers are beginning to explore the effects. Will environmental threats to our oceans trigger a new wave of anxiety disorders? Eco-anxiety is now recognized as condition by the American Psychological Association [13]. Some argue that it is not a mental disorder, but a rational response to current conditions. In either case, clinicians are increasingly encountering patients and attempting to treat this condition. Eco-anxiety is a maladaptive response to increasing levels of pollution, ecological disasters, and threats to the natural environment. While much of this can be triggered by the relentless alarms present in the media, visual reminders of increasing pollution play a significant role. The oceans

5  The Blue Mind

63

were once thought to be so vast that humankind would have little ability to affect its pristine nature. No doubt this concept has been part of the spiritual construct that literature and the arts have referred to, a place we can retreat to physically, mentally, and spiritually because it is separate from the activities of man. Is that emotional refuge in jeopardy now? Plastics in significant amounts are found in all parts of the marine world, from the remotest islands to the artic and in deepest area known, the Marianas Trench. Photographs from around the world display plastic bottles and grocery bags washing ashore and sea animals getting suffocated by fishing nets, a stark reminder that the space on earth is limited and human’s footprint is large. Plastic micro-beads found in everyday products inevitably find their way into the ocean and enter the food chain when sea animals consume them. In human beings and animals, ingestion of plastic leads to detrimental health outcomes: internal bleeding, abrasions, ulcers, and intestinal tract blockages. [C] Ultimately, the earth’s bioaccumulation of plastic highlights its adverse effects on environmental and human health [14]. Despite transitions to non-fossil fuels, oil spills will continue. Oil spills can have a marked emotional impact due to their direct visual impact on wildlife, vegetation, and coastal spaces. The Deepwater Horizon oil spill inundated the coast with blankets of black petroleum that coated vegetation wildlife and induced fish kills. Many in that region have vivid images of tar balls that washed up on distant beaches for months [15]. The definitive solution is to reverse and limit human impact on the oceans, and this should continue to be our objective. Discernibly, this solution is not realistic in the near future, and conditions will probably worsen before they hopefully improve. In the interim, it will be important to understand the impact of the blue environment on mental and physical health and take steps to maximize and preserve its benefits. As the world becomes even more connected, blue spaces will become more valuable. As is often the case, economics may determine who has access to those areas. Policy that allows for fair allocation of those spaces and protects the rights of all will be needed.

5.4  The Need to Communicate Success Stories Much is heard about the threats to the oceans. As expected, it is taking its mental toll. We run the risk developing an overwhelming sense of doom that paralyzes us from action. At this point, an equal amount of communication should be undertaken to show success stories and future positive attempts to improve aquatic environments. Efforts such as the United Nations Global Initiative and the newly described “Blue New Deal” are invaluable toward this goal. The themes in these efforts are global and ambitious. San Francisco, California, set a precedent by becoming the first major US city to ban certain plastics in July 2018 [16]. Simply showcasing that we are making progress in an organized and deliberate fashion can be a driver of individual efforts and a chance to energize differing environmental groups.

64

K. Conrad et al.

As it is often the case, when a source of our mental well-being is threatened, we finally come to understand its value. The ecological threat to our oceans is not abstract in the slightest, but rather one that can be traced directly to human health. Worse, the ecological threat is largely our own doing. Every effort should be made to research and communicate these direct impacts, before the issues overwhelm us and a sense of crippling futility sets in.

References 1. Honda, T., & Kamioka, H. (2012). Curative and health enhancement effects of aquatic exercise: Evidence based on interventional studies. Open access Journal of Sports Medicine, 3, 27–34. Published 29 March 2012. https://doi.org/10.2147/OAJSM.S30429. 2. Nichols, W. (2104). Blue mind: The surprising science that shows how being near, in, on, or under water can make you happier, healthier, more connected, and better at what you do. New York: Little Brown and Company. 3. Collins, F.  S., Fleming, R., Rutter, D., Iyengar, S., Tottenham, N., Patel, A.  D., Limb, C., Johnson, J. K., & Holochwost, S. J. (2018). NIH/Kennedy Center Workshop on music and the brain: Finding harmony. Neuron, 97(6), 1214–1218. 4. Nutsford, D., Pearson, A. L., Kingham, S., & Reitsma, F. (2016). Residential exposure to visible blue space (but not green space) associated with lower psychological distress in a capital city. Health Place, 39, 70. https://doi.org/10.1016/j.healthplace.2016.03.002. 5. Hignett, A., White, M. P., Pahl, S., Jenkin, R., & Froy, M. L. (2018). Evaluation of a surfing programme designed to increase personal well-being and connectedness to the natural environment among ‘at risk’ young people. Journal of Adventure Education and Outdoor Learning, 18, 53. https://www.tandfonline.com/doi/abs/10.1080/14729679.2017.1326829. 6. Crouse, D. L., Balram, A., Hystad, P., Pinault, L., van den Bosch, M., Chen, H., Rainham, D., Thomson, E. M., Close, C. H., van Donkelaar, A., Martin, R. V., Ménard, R., Robichaud, A., & Villeneuve, P.  J. (2018). Associations between living near water and risk of mortality among urban Canadians. Environmental Health Perspectives, 126(7), 077008. https://doi. org/10.1289/EHP3397. eCollection 2018 July. 7. Dzhambov, A.  M. (2018). Residential green and blue space associated with better mental health: a pilot follow-up study in university students. Arh Hig Rada Toksikol, 69(4), 340–349. https://doi.org/10.2478/aiht-2018-69-3166. 8. Pearson, A. L., Shortridge, A., Delamater, P. L., Horton, T. H., Dahlin, K., Rzotkiewicz, A., et al. (2019). Effects of freshwater blue spaces may be beneficial for mental health: A first, ecological study in the North American Great Lakes region. PLoS One, 14(8), e0221977. 9. Garrett, J. K., Clitherow, T. J., White, M. P., Wheeler, B. W., Fleming, L. E., Oja, P., Kelly, P., Pedisic, Z., et al. (2019). Coastal proximity and mental health among urban adults in England: The moderating effect of household income. Health & Place, 59, 102200. 10. https://search-proquest-com.ezproxy.library.uq.edu.au/docview/1730204593?rfr_ id=info%3Axri%2Fsid% 3Aprimo. 11. Oja, P. (2017). Associations of specific types of sports and exercise with all-cause and cardiovascular-­disease mortality: A cohort study of 80 306 British adults. British Journal of Sports Medicine, 51, 812–817. 12. www.wfmz.com/news/cnn-national/gulf-coast - b e a c h e s- c l o se d - d u e - t o - a l ga e bloom/1090442825. Retrieved 1 November 2019. 13. Whitmore-Williams, S. C., Manning, C., Krygsman, K., & Speiser, M. (2017). Mental health and our changing climate (PDF). Washington, DC: American Psychological Association. 14. https://www-ncdi-nim-nih-gog.exproxy.;ibray.yq.edu.au/pubmed/31709817

5  The Blue Mind

65

15. Norse, E. A., & Amos, J. (2010). Impacts, perception, and policy implications of the BP/deepwater horizon oil and gas disaster (PDF). Environmental Law Reporter, 40(11), 11058–11073. ISSN 0046-2284. Retrieved 1 November 2019. 16. https://search-proquest-com.ezproxy.library.uq.edu.au/docview/2140861743?rfr_ id=info%3Axri%2Fsid% 3Aprimo.

Chapter 6

Oceans and Rapid Climate Change Wei Liu and Alexey Fedorov

Oceans can play a vital role in abrupt climate change. Specifically, rapid variations in the Atlantic Meridional Overturning Circulation (AMOC), one the major ocean current systems, are believed to be the primary driver of past and potentially future abrupt climate changes. The AMOC is driven by a combination of large-scale ocean density gradients, which depend on temperature and salinity contrasts, and wind forcing. In the upper North Atlantic Ocean, warm water is transported northward by the Gulf Stream and then the North Atlantic Current. This water becomes progressively colder and more saline along its path due to surface heat loss and evaporation; when it reaches high latitudes of the North Atlantic, it is dense enough to sink deep into the ocean. This deep dense water, referred to as North Atlantic Deep Water (NADW), slowly spreads southward several kilometers below the surface. Zonally integrating the components of surface and deep currents over the Atlantic Ocean yields the characteristic AMOC structure on the depth-latitude plane with an upper limb of northward near-surface warm flow and a lower limb of denser, colder southward return flow (Fig. 6.1). The AMOC is a part of the global ocean “convey belt” that regulates oceanic transports of heat, freshwater, and carbon [1]. Under modern climate conditions, the AMOC has a strength of about 17 Sv (1 Sv =106 m3/s) as measured by the Rapid Climate Change-Meridional Overturning Circulation and Heat Flux Array (RAPID-­ MOCHA) across ~26°N since 2004 [2] (Fig. 6.2). This energetic ocean circulation carries tremendous oceanic heat northward and thus contributes to the moderate climate of the British Isles and northwest Europe [3]. Should the AMOC collapse, W. Liu (*) Department of Earth and Planetary Sciences, University of California Riverside, Riverside, CA, USA e-mail: [email protected] A. Fedorov Department of Geology and Geophysics, Yale University, New Haven, CT, USA LOCEAN/IPSL, Sorbonne University, Paris, France © Springer Nature Switzerland AG 2021 K. Conrad (ed.), From Hurricanes to Epidemics, Global Perspectives on Health Geography, https://doi.org/10.1007/978-3-030-55012-7_6

67

68

W. Liu and A. Fedorov

Fig. 6.1  Atlantic meridional overturning stream function (units: Sverdrup or Sv = 106 m3/s) based on the annual mean data of the ECCO (Estimating the Circulation and Climate of the Ocean) dataset during 1992–2015 [73]. Black arrows indicate the circulation direction of the AMOC

Fig. 6.2  A temperature-based AMOC index (blue) reconstructed from the NASA GISS surface temperature data [74] based on the method by Ramhstorf et al. [75], and the AMOC strength from the RAPID-MOCHA observations (red, c.f. [2])

6  Oceans and Rapid Climate Change

69

its poleward heat transport would decrease, causing a significant cooling over the North Atlantic and surrounding regions, as was dramatically depicted in the movie “The Day after Tomorrow,” albeit in a very exaggerated fashion. In addition to surface air temperature change, a collapsed AMOC would increase storminess and episodes of extreme weather over Europe [4], shift the tropical rain belt southward [5] affecting precipitation over the Sahel region of Africa, weaken the North Atlantic ocean sink for carbon dioxide [6], and affect the hurricanes [7] and marine ecosystems [8] in the North Atlantic and even the El Nino-Southern oscillation (ENSO) in the tropical Pacific [9]. Given the large potential climate impacts of AMOC changes, it is of public concern whether the AMOC could indeed collapse in the future as a result of anthropogenic climate change. To answer this question, we will first review available paleo-evidence, focusing on past AMOC variations and the associated climate response, and explore what we can learn from the past. We then look at the stability and future projections for the AMOC.

6.1  Insights from Past Climates Paleo-records contain strong evidence of abrupt climate changes such as the warm Dansgaard-Oeschger (D/O) [10] and cold Heinrich events [11, 12] that occurred on millennial timescales during the last glacial cycle (Fig. 6.3). These events as well as the Younger Dryas cooling that punctuated the last deglaciation are considered examples of abrupt climate changes as their onset typically occurred over the course of several decades [13] and they brought about large temperature changes on the order of 10 °C or more to North Atlantic. D/O and Heinrich events are generally characterized by opposite temperature changes in the Northern and Southern Hemispheres consistent with the so-called bipolar “see-saw” pattern [14]. Particularly, Greenland ice-core records reveal abrupt cooling of 10–16 °C between cold (Greenland stadial) and warmer (Greenland interstadial) climate states [15], while Antarctic ice cores exhibit temperature variations of opposite sign [16, 17]. Following these inter-hemispheric temperature changes, the tropical rain belt of the atmospheric Intertropical Convergence Zone (ITCZ) migrated meridionally toward warmer sea surface temperatures [18, 19]. These past abrupt climate changes are believed to have been intrinsically linked to rapid changes in the AMOC (e.g., [14, 20, 21]), or more fundamentally with the stability of the AMOC. For example, the absence of rapid climate changes during the Holocene suggests that the AMOC has remained in a stable regime, possibly a mono-stable regime (only having one equilibrium state), in the last ~10,000 years [22] (Fig. 6.3). On the other hand, the switch between interstadial and stadial modes in the D/O cycles is hypothesized to indicate the transition between different equilibrium states of a bi-stable AMOC [23, 24] or alternatively a chaotic transition between different AMOC unstable states (e.g., [25, 26]). The hypothesis linking past abrupt climate changes to AMOC variations is supported by the paleo-evidence

70

W. Liu and A. Fedorov

Fig. 6.3  Evidence for a weaker AMOC during Heinrich and the Younger Drays events (marked as gray bands). Oxygen isotope ratio in the planktonic foraminifer Neogloboquadrina pachyderma from the western North Atlantic [76] is shown in green; low values reflect the presence of glacial meltwater. 231Pa/230Th ratio measured in sediments on the Bermuda Rise is shown in purple [29], plum [77], and magenta [78], which reflects changes in the residence time of deep water in the Atlantic potentially related to changes in the AMOC—higher values mean a weaker AMOC. Note that the vertical axis is inverted for the top and middles curves. Mean 6C13 values for the deep Atlantic (an average between several records) are shown in blue [28]; low values indicate the prevalence of southern source waters, rather than northern source water, which may indicate AMOC weakening. Additional abbreviations: kyr BP 1000 years before present; HS Heinrich stadial; YD Younger Dryas. Note that there are uncertainties on the dating of these events in different time series

6  Oceans and Rapid Climate Change

71

of reduced NADW production from the interstadial mode to the stadial mode during a D/O event [27, 28] or a longer residence time of deep water in the North Atlantic during Heinrich events [29], which indicates a weakened or collapsed AMOC during the events (Fig. 6.3). One important issue in the relationship between abrupt climate changes and AMOC variations is the necessity for trigger mechanisms of the transitions between AMOC equilibrium states (if they are indeed stable). For example, the Meltwater Pulse 1A (MWP-1A) has been suggested as a mechanism to trigger abrupt AMOC changes and thereby the transition from the Heinrich Event 1 to Bølling-Allerød warming. However, this meltwater mechanism is challenged by the uncertainties in the lead-lag relationship between meltwater pulse and AMOC change. The underlying reason lies in the poor chronology of paleo-reconstructions, which makes it difficult to achieve precise timing and location of meltwater pulses (such as MWP-1A) [30] and therefore to establish a precise chronological order between meltwater pulses and AMOC changes. Despite uncertainties in the triggering mechanisms, and even though the hypothesis of an unstable AMOC during the last glacial interval cannot be excluded, many past abrupt climate events are still considered most likely to be related to a bi-stable AMOC.  These insights from past climate changes raise the question whether abrupt AMOC changes can occur in the near future.

6.2  The Stability of the AMOC The RAPID-MOCHA observations suggest that the AMOC has been slowing down during the past decade ([3] and Fig. 6.2). Although the robustness of this decadal decline of the AMOC remains unclear in view of natural variability in the climate system, on longer timescales temperature-based proxies of AMOC strength also suggest a gradual AMOC weakening since the mid-twentieth century (e.g., [31]). Could these observations indicate the possibility of a rapid AMOC weakening or even collapse in the near future? This risk, however, has been downplayed by the climate model projections of the latest Intergovernmental Panel on Climate Change (IPCC) Fifth Assessment Report [32] who concluded that it would be “extremely unlikely” that the AMOC collapses in the twenty-first century (meaning that the probability of the AMOC collapse is less than 10%). Nevertheless, the reliability of these climate projections is challenged by the ubiquitous biases present in climate models. To achieve a creditable climate projection, we need to bridge the gap between the observations and climate simulations, which in turn requires a deeper understanding of AMOC multiple equilibria. The characteristics of the AMOC multiple equilibria result from the nonlinear nature of the circulation. Earlier box-model studies suggest that the positive salinity-­ advection feedback destabilizes the AMOC, leading to a bi-stable circulation [33– 35]. The existence of AMOC multiple equilibria has been demonstrated with different box models, loop models of the ocean flow [26, 36], models with two

72

W. Liu and A. Fedorov

connected basins [37], two-dimensional ocean models [38], ocean general circulation models [39–42], earth system models of intermediate complexity (EMICs, c.f. [24, 31, 43]), and some coupled climate models [44–47]. In contrast to the bi-stable AMOC often observed in simplified models, such a regime does not seem to be a robust feature across comprehensive coupled climate models. For example, Stouffer et  al. tested the AMOC stability in several models from the Coupled Model Intercomparison Project/Paleo-Modeling Intercomparison Project (CMIP/PMIP) but found no evidence of AMOC multiple equilibria except for one model [48]. The systematic lack of AMOC multiple equilibria across coupled climate general circulation models (GCMs) points to several factors that tend to over-stabilize the AMOC, including oceanic diffusivity and mixing [49, 50], atmospheric internal variability [51, 52], wind stress feedbacks [38, 53, 54], and ocean-atmosphere coupling [45]. However, each of these factors has complex and sometimes counterintuitive or conflicting effects on AMOC multiple equilibria. For example, increasing oceanic diapycnal diffusivity has been suggested either to suppress AMOC multiple equilibria by generating a more diffusive and linear circulation by Manabe and Stouffer [49] or to enhance AMOC multiple equilibria via a stronger oceanic upwelling in Prange et al. [50]. Wind stress feedbacks have been argued to stabilize the AMOC in the modern climate [53] but to destabilize the AMOC under glacial conditions [54]. To evaluate the stability of the AMOC offline in a comprehensive climate model, and more importantly in nature using available observations, one has to formulate a diagnostic indicator of AMOC stability [55, 56, 57]. Based on the Atlantic-Arctic freshwater budget [56], Rahmstorf [55] first proposed the freshwater export by the ∗ AMOC across the southern boundary (~34°S) of the Atlantic (FOT or M ovS ) as a diagnostic indicator of the AMOC stability. This freshwater export indicator was ∗ ∗ ∗ ∗ later refined as a divergence indicator ( ∆M ov ) where ∆M ov  =  M ovS  −  M ovN [58– ∗ 60] by including freshwater transport across the northern boundary ( M ovN ) to consider the effect coming from the Arctic. The physics behind this indicator is related to the basin-scale salinity-advection feedback [33]. Suppose an active AMOC produces a climate state with an Atlantic ∗     0), an initial AMOC weakening will reduce freshwater convergence and increase the salinity of the Atlantic, facilitating deep convection and preventing the further weakening of the AMOC. In this case, the AMOC tends to recover to its original state, and thus it is in a mono-stable regime. ∗ As a powerful metric, ∆M ov can be diagnosed from the observations and used to estimate the stability of a realistic AMOC under modern climate conditions, helping

6  Oceans and Rapid Climate Change

73

to explain the systematic lack of AMOC multiple equilibria in coupled climate ∗ GCMs. Liu et al. [61] estimated the values of ∆M ov from the available observations and used them as the baseline to assess the AMOC stability in coupled GCMs. Their synthesis of Atlantic freshwater transports from ship, buoy, and hydrographic obser∗ ∗ vations showed the range of M ovS between −0.34 and −0.1 Sv and M ovN of about ∗ 0.15 Sv. This yields the observed range for ∆M ov of −0.19 to +0.05 Sv. Liu et al. ∗ [63] further diagnosed ∆M ov from multiple oceanic reanalysis products and found consistently negative values for this stability indicator. Thus, both instrumental observations and reanalysis data suggest an AMOC induced freshwater divergence ∗ (i.e.. ∆M ov   0), implying a mono-stable AMOC. This mono-stable AMOC in coupled climate models offers a possible explanation for the rapid recovery of the AMOC in freshwater perturbation experiments [48] and suggests a critical bias across coupled climate models in AMOC stability. Liu et al. [61–63] further find that the bias in ∗ AMOC stability (i.e., a positive ∆M ov ) is mostly caused by a biased freshwater ∗ transport through the south boundary. At the northern boundary M ovN is consistently negative in both observations and models as associated with freshwater import from the Arctic, which is also important in reversing the sign of freshwater divergence in some models. However, at the southern boundary, observations suggest a strong freshwater export due to a more saline northward flow of surface and thermocline waters relative to the deep southward flow of NADW, whereas coupled climate models show a freshwater import or diminished freshwater transport primarily due to an upper-layer fresh bias. The systematic biases in salinity and hence in freshwater transport and divergence has been argued to originate, at least partly, from the persistent tropical bias due to the common double ITCZ problem [61, 62]. Coupled GCMs typically suffer from this bias in simulated annual mean climate, which is characterized by two precipitation bands, one on each side of the equator, and an excessively strong cold tongue that penetrates too far westward along the equator [64, 65]. This double ITCZ bias is even more extreme in the tropical Atlantic, causing excessive rainfall over the tropical Atlantic south of the equator, which freshens and then creates a negative salinity bias over the tropical South Atlantic. As a result, the AMOC upper branch across the southern boundary of the Atlantic transports water which is too fresh, so that the total southward freshwater export diminishes or even reverses, turning into freshwater import. This contributes to the Atlantic freshwater conver∗   >  0) and likely a mono-stable AMOC, across many coupled cligence ( ∆M ov mate GCMs. In an effort to validate the AMOC stability criterion and to explore the cause of the AMOC stability bias, Liu et  al. [61] corrected the AMOC stability bias in a broadly used coupled climate GCM, CCSM3 developed by NCAR, by applying surface flux adjustments. As other coupled climate GCMs, the original version of NCAR CCSM3 shows a fresh bias in the upper ocean at 34°S, an Atlantic freshwa∗ ter convergence ( ∆M ov  > 0) and, indeed, a mono-stable AMOC in modern climate.

74

W. Liu and A. Fedorov

Starting from this original model, Liu et  al. [61] built a corrected model using a surface flux adjustment method. The addition of heat and freshwater fluxes ­eliminated most of temperature and salinity biases at the ocean surface and, more importantly, corrected salinity bias in the upper ocean in the South Atlantic. This corrected version of NCAR CCSM3 successfully simulates a strong freshwater ∗ export across the southern boundary, an Atlantic freshwater divergence ( ∆M ov