Sustainable Built Environments [2nd ed.] 9781071606834, 9781071606841

880 136 37MB

English Pages [717] Year 2020

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Sustainable Built Environments [2nd ed.]
 9781071606834, 9781071606841

Citation preview

Encyclopedia of Sustainability Science and Technology Series Editor-in-Chief: Robert A. Meyers

Vivian Loftness  Editor

Sustainable Built Environments Second Edition A Volume in the Encyclopedia of Sustainability Science and Technology, Second Edition

Encyclopedia of Sustainability Science and Technology Series Editor-in-Chief Robert A. Meyers

The Encyclopedia of Sustainability Science and Technology series (ESST) addresses the grand challenge for science and engineering today. It provides unprecedented, peer-reviewed coverage in more than 600 separate articles comprising 20 topical volumes, incorporating many updates from the first edition as well as new articles. ESST establishes a foundation for the many sustainability and policy evaluations being performed in institutions worldwide. An indispensable resource for scientists and engineers in developing new technologies and for applying existing technologies to sustainability, the Encyclopedia of Sustainability Science and Technology series is presented at the university and professional level needed for scientists, engineers, and their students to support real progress in sustainability science and technology. Although the emphasis is on science and technology rather than policy, the Encyclopedia of Sustainability Science and Technology series is also a comprehensive and authoritative resource for policy makers who want to understand the scope of research and development and how these bottom-up innovations map on to the sustainability challenge. More information about this series at http://www.springer.com/series/15436

Vivian Loftness Editor

Sustainable Built Environments Second Edition A Volume in the Encyclopedia of Sustainability Science and Technology, Second Edition

With 388 Figures and 57 Tables

Editor Vivian Loftness Center for Building Performance and Diagnostics Carnegie Mellon University Pittsburgh, PA, USA

ISBN 978-1-0716-0683-4 ISBN 978-1-0716-0684-1 (eBook) ISBN 978-1-0716-0685-8 (print and electronic bundle) https://doi.org/10.1007/978-1-0716-0684-1 1st edition: © Springer Science+Business Media New York 2013 2nd edition: © Springer Science+Business Media, LLC, part of Springer Nature 2020 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 Science+Business Media, LLC, part of Springer Nature. The registered company address is: 1 New York Plaza, New York, NY 10004, U.S.A.

There are dozens who I would and should dedicate this volume to, from my partner in work and life, Volker, to my always collaborative colleagues, to my always supportive kids, to the sustainability leaders who have authored the chapters in this Encyclopedia. Yet, I really need to dedicate this volume to my father, Robert L. Loftness, who inspired me to be a scientist and an environmentalist and a writer and an architect and so much more. He wrote his own encyclopedia, the Energy Handbook, while leading the Washington, DC office of EPRI. He kept files of critical research on every possible source of energy and energy efficiency, and he welcomed my debates on the matched relevance of efficiencies in energy demand in the built environment and innovations in supply. Moreover, he was the gentlest, smartest, most loving father one could ever have. To Dad.

Series Preface

Our nearly 1000-member team recognizes that all elements of sustainability science and technology continue to advance as does our understanding of the needs for energy, water, clean air, food, mobility, and health and the relation of every single aspect of this vast and interconnected body of knowledge to climate change. Our Encyclopedia content is at a level for university students, professors, engineers, and other practicing professionals. It is gratifying for our team to note that our online First Edition has been heavily utilized as evidenced by over 500,000 downloads which of course is in addition to scientists’ utilization of the Encyclopedia and individual “spin-off” volumes in print. Now we are pleased to have a Living Reference online which assures the sustainability community that we are providing the latest peer-reviewed content covering the science and technology of the sustainability of the Earth. We are also publishing the content as a Series of individual topical books for ease of use by those with an interest in particular subjects, and with expert oversight in each field to ensure that the Second Edition presents the state-of-the-science today. Our team covers the physical, chemical, and biological processes that underlie the Earth system including pollution and remediation and climate change, and we comprehensively cover every energy and environment technology as well as all types of food production, water, transportation, and the sustainable built environment. Our team of 15 board members includes two Nobel Prize winners (Kroto and Fischlin), two former Directors of the NSF (Colwell and Killeen), the former President of the Royal Society (Lord May), and the Chief Scientist of the Rocky Mountain Institute (Amory Lovins). And our more than 40 eminent section editors and now book editors assure quality of our selected authors and their review presentations. The extent of our coverage clearly sets our project apart from other publications which now exist, both in extent and depth. In fact, current compendia of the science and technology of several of these topics do not presently exist and yet the content is crucial to any evaluation and planning for the sustainability of the Earth. It is important to note that the emphasis of our project is on science and technology and not on policy and positions. Rather, policy makers will use our presentations to evaluate sustainability options. Vital scientific issues include: human and animal ecological support sys tems; energy supply and effects; the planet’s climate system; systems of

vii

viii

Series Preface

agriculture, industry, forestry, and fisheries and the ocean; fresh water and human communities; waste disposal; transportation and the built environment in general and the various systems on which they depend, and the balance of all of these with sustainability. In this context, sustainability is a characteristic of a process or state that can be maintained at a certain level indefinitely even as global population increases toward 9 billion by 2050. The population growth, and the hope for increase in wealth, implies something like a 50% increase in food demand by as early as 2030. At the same time, the proportion of the population that lives in an urban environment will go up from about 47% to 60%. Global economic activity is expected to grow 500%, and global energy and materials use is expected to increase by 300% over this period. That means there are going to be some real problems for energy, agriculture, and water, and it is increasingly clear that conflicting demands among biofuels, food crops, and environmental protection will be difficult to reconcile. The “green revolution” was heavily dependent on fertilizers which are manufactured using increasingly expensive and diminishing reserves of fossil fuels. In addition, about 70% of available freshwater is used for agriculture. Clearly, many natural resources will either become depleted or scarce relative to population. Larkspur, CA, USA July 2020

Robert A. Meyers, Ph.D. Editor-in-Chief

Volume Preface

Despite the immediate challenges of a global pandemic, climate change remains the existential threat of our century. We have completely tipped the balance of ecosystems with extraction, consumption, and waste – with global warming, climate disruptions, air and water pollution, and pandemics as the consequences. Many believe the answer is in renewable sources that will enable us to continue consumption at the levels we have grown accustomed, despite serious global inequities in that consumption. They believe renewable sources will allow the voracious demands in the building, transportation, and industrial sectors to continue and even expand worldwide. This is not a vision for a sustainable, ecologically balanced, and abundant future. Instead, we need to discover innovative ways to live within the cycles of nature, enjoying the diversity and dynamics of place. This is the fundamental tenant of sustainable built environments. This Encyclopedia has gathered 25 leading authors from around the world to provide insight into the breadth of knowledge needed to fully define and understand sustainable built environments. The chapters are grouped in four major parts and ordered to build expertise across disciplines and scales of practice. For architects and designers as well as building construction and facilities managers, new models and theories are critical to rebuilding the ecological balance that is our planet. Five models of sustainable built environments are introduced – Bioclimatic Design, Biophilic Design, Healthy Built Environments, Regenerative Design, and Resilient Design – with the sixth model Cradle to Cradle Design captured in the last chapter of this Encyclopedia. Each of these authors realize that the word sustainable is inadequate to describe the rebuilding of ecological balance and abundance, and offer addenda to an already rich set of goals captured in the green building movement worldwide. After decades of focusing on sustainable buildings, the environmental design community realized that design decision-making at the urban scale was even more critical to rebalancing our ecosystem. It is critical to repair our land use, soils, farms, and forests to reverse the disruption – mitigating climate change through carbon sequestration and biodiversity building, and adapting to the repercussions of climate change – floods, droughts, hurricanes, fires, and more. Six leading practitioners offer expertise in urban design for sustainability with chapters spanning from Sustainable Urbanism, Water and Sustainable

ix

x

Volume Preface

Design, Green Infrastructures to Face Climate Change, and Greening Brownfields to Biodiversity in Cities and Biophilic Cities. While sustainable urbanism is a critical precursor, buildings must also be designed and renovated to become a critical partner in the ecological balance. Nine leading researchers and practitioners provide deep expertise on Building Design for Sustainability, beginning with building enclosure design to maximize natural conditioning and minimize energy demand with successively enriched design detailing of Facades and Enclosures, Passive House (deep conservation), Passive Solar Heating, Daylighting, and Natural Ventilation. With environmental conditioning loads reduced to 20–40% of their former levels, the design of mechanical systems can be dramatically improved as explained in the two chapters on Sustainable Heating Ventilation and Air Conditioning and Adaptive Comfort and Mixed-Mode Conditioning. While a number of important entries should be added to address electric loads and renewable energy generation at the building level, other volumes of the Springer Encyclopedia on Sustainability help to fill this gap. Two additional chapters in this part capture the potential of the fifth fac¸ade – the roof for carbon capture and UHI reduction, and the sixth fac¸ade – foundations for geothermal energy, towards creating a low energy, environmentally balanced future. The complexity and richness of design for sustainability must span the full range of building scales from material to region, of building disciplines from specifier to urban designer, and of building climates from drought to deluge, tropic to arctic. This requires new forms of project delivery that celebrate integrated design thinking. Five leading practitioners and researchers define Integrated Delivery Processes, Rating Systems for Sustainability, Sustainability Performance Simulation Tools for Building Design, Evidence-Based Design, and the Role of Buildings in Materials Banking and Resource Repletion. This closing chapter on Materials Passports raises the question of whether every new or renovated project in the built environment should have a passport capturing its contribution to a regenerative built environment at the material, building, urban, and regional scale. For the moment, this is the definitive encyclopedia on sustainability in the built environment, written by world leaders in each subject area. These authors have written the textbooks, the codes and standards, and the vision for a more sustainable built environment, at every level – regional, urban, community, building, and material. They have transformed practice, research, and education in fundamental ways and it has been a privilege to work with them to assemble this 2020 Encyclopedia of Sustainable Built Environments. Pittsburgh, PA, USA September 2020

Vivian Loftness, FAIA Volume Editor

Contents

Sustainable Built Environments: Introduction . . . . . . . . . . . . . . . . Vivian Loftness Part I

1

Models of Sustainable Built Environments . . . . . . . . . . . .

17

Bioclimatic Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Donald Watson

19

Biophilic Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Catherine O. Ryan and William D. Browning

43

Sustainable and Healthy Built Environment . . . . . . . . . . . . . . . . . . Vivian Loftness and Megan Snyder

87

Regenerative Development and Design . . . . . . . . . . . . . . . . . . . . . . 115 Pamela Mang and Bill Reed Resilient Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143 Alex Wilson and Mary Ann Lazarus Part II

Urban Design for Sustainability . . . . . . . . . . . . . . . . . . . . . .

161

Sustainable Urbanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163 Douglas Farr Water and Sustainable Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187 Herbert Dreiseitl Green Infrastructures to Face Climate Change in an Urbanizing World . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207 Stephan Pauleit, Ole Fryd, Antje Backhaus, and Marina Bergen Jensen Greening Brownfields in Urban Redevelopment . . . . . . . . . . . . . . . 235 Juliane Mathey and Dieter Rink Biodiversity in Cities, Reconnecting Humans with Nature . . . . . . . 251 Robbert P. H. Snep and Philippe Clergeau

xi

xii

Contents

Biophilic Cities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275 Timothy Beatley Part III Building Design for Sustainability . . . . . . . . . . . . . . . . . . .

293

Facades and Enclosures: Building for Sustainability . . . . . . . . . . . 295 Volker Hartkopf, Azizan Aziz and Vivian Loftness Passive House (Passivhaus) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 327 Katrin Klingenberg Passive Solar Heating in the Built Environment . . . . . . . . . . . . . . . 351 Robert Hastings Daylighting Controls, Performance, and Global Impacts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 383 Helmut Köster Natural Ventilation in Built Environment . . . . . . . . . . . . . . . . . . . . 431 Tong Yang and Derek J. Clements-Croome Sustainable Heating Ventilation and Air Conditioning . . . . . . . . . . 465 Kevin Hydes and Jennifer Fosket Adaptive Comfort and Mixed-Mode Conditioning . . . . . . . . . . . . . 481 Richard de Dear and Gail Brager Green Roofs: Ecological Functions of the Fifth Facade . . . . . . . . . 495 Manfred Köhler and Andrew Michael Clements Geothermal Conditioning: Critical Sources for Sustainability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 519 Nina J. Baird Part IV

Sustainability Practices . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

549

Sustainable Design and Construction, Integrated Delivery Processes, and Building Information Modeling . . . . . . . . . . . . . . . . 551 Laura Lesniewski and Bob Berkebile Rating Systems for Sustainability . . . . . . . . . . . . . . . . . . . . . . . . . . . 573 Raymond J. Cole Sustainability Performance Simulation Tools for Building Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 589 Khee Poh Lam Evidence-Based Design for Indoor Environmental Quality and Health . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 657 Charlene W. Bayer

Contents

xiii

Materials Banking and Resource Repletion, Role of Buildings, and Materials Passports . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 677 Katja Hansen, Michael Braungart and Douglas Mulhall Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 703

About the Editor-in-Chief

Robert A. Meyers President: RAMTECH Limited Manger, Chemical Process Technology, TRW Inc. Post-doctoral Fellow: California Institute of Technology Ph.D. Chemistry, University of California at Los Angeles B.A., Chemistry, California State University, San Diego

Biography Dr. Meyers has worked with more than 20 Nobel laureates during his career and is the originator and serves as Editor in Chief of both the Springer Nature Encyclopedia of Sustainability Science and Technology and the related and supportive Springer Nature Encyclopedia of Complexity and Systems Science.

Education Postdoctoral Fellow: California Institute of Technology Ph.D. in Organic Chemistry, University of California at Los Angeles B.A., Chemistry with minor in Mathematics, California State University, San Diego Dr. Meyers holds more than 20 patents and is the author or Editor in Chief of 12 technical books including the Handbook of Chemical Production Processes, Handbook of Synfuels Technology, and Handbook of Petroleum xv

xvi

Refining Processes now in 4th Edition, and the Handbook of Petrochemical Production Processes, now in its second edition, (McGraw-Hill) and the Handbook of Energy Technology and Economics, published by John Wiley & Sons; Coal Structure, published by Academic Press; and Coal Desulfurization as well as the Coal Handbook published by Marcel Dekker. He served as chairman of the Advisory Board for A Guide to Nuclear Power Technology, published by John Wiley & Sons, which won the Association of American Publishers Award as the best book in technology and engineering.

About the Editor-in-Chief

About the Editor

Vivian Loftness, FAIA, is an internationally renowned researcher, author, and educator focused on environmental design and sustainability, climate and regionalism in architecture, and the integration of advanced building systems for health and productivity. At Carnegie Mellon University, Professor Loftness holds the Paul Mellon Distinguished Chair in Architecture, is one of 40 University Professors, and served a decade as Head of the School of Architecture. With over 30 years of industry and government research funding, she is a key member of Carnegie Mellon’s leadership in sustainability research and education and contributor to the ongoing development of the Intelligent Workplace – a living laboratory of commercial building innovations for performance. Her collaborative research is captured in over 100 journal articles, book chapters, and books, as well in the 2013 and 2020 Springer Reference Encyclopedia of Sustainable Built Environments, for which she serves as Editor. Professor Loftness has served on over 25 Boards of Directors, including EPA’s NACEPT, DOE’s FEMAC, and the National USGBC, AIA, and ILFI Boards. She has served on 12 National Academy of Science panels as well as the Academy’s Board on Infrastructure and the Constructed Environment and given four Congressional testimonies on sustainable design. Her work has influenced national policy and building projects, including the Adaptable Workplace Lab at the U.S. General Services Administration and the Laboratory for Cognition at Electricity de France. xvii

xviii

Professor Loftness has been recognized as a LEED Fellow; a Senior Fellow of the Design Futures Council, the New Buildings Institute, and the Scott Energy Institute; and one of 13 Stars of Building Science by the Building Research Establishment in the UK. She received Awards of Distinction from AIA Pennsylvania and NESEA, holds a National Educator Honor Award from the American Institute of Architecture Students, and a “Sacred Tree” Award from the US Green Building Council. Professor Loftness is a Bachelor of Science and Master of Architecture from MIT.

About the Editor

Contributors

Azizan Aziz Center for Building Performance and Diagnostics, Carnegie Mellon University, Pittsburgh, PA, USA Antje Backhaus gruppe F, Berlin, Germany Nina J. Baird Center for Building Performance and Diagnostics, Carnegie Mellon University, Pittsburgh, PA, USA Charlene W. Bayer Hygieia Sciences LLC, Johns Creek, GA, USA Timothy Beatley School of Architecture, University of Virginia, Charlottesville, VA, USA Bob Berkebile FAIA, BNIM, Kansas City, MO, USA Gail Brager University of California, Berkeley, CA, USA Michael Braungart EPEA Internationale Umweltforschung GmbH, Hamburg, Germany Institute for Ethical and Transdisciplinary Sustainability, Leuphana University, Lüneburg, Germany William D. Browning Terrapin Bright Green, New York, NY, USA Andrew Michael Clements Oikosteges Oikosystem Roofs, Northampton, UK Derek J. Clements-Croome School of Construction Management and Engineering, University of Reading, Reading, UK School of Engineering and Materials Science, Queen Mary University of London, London, UK Philippe Clergeau Department of Ecology and Biodiversity Management, National Museum of Natural History, Paris, France Raymond J. Cole School of Architecture and Landscape Architecture, University of British Columbia, Vancouver, BC, Canada Richard de Dear The University of Sydney, Sydney, NSW, Australia Herbert Dreiseitl DREISEITL consulting GmbH, Überlingen, Germany Douglas Farr Founding Principal, Farr Associates, Chicago, IL, USA xix

xx

Jennifer Fosket Social Green, Berkeley, CA, USA Ole Fryd Landscape Architecture and Planning, Department of Geosciences and Natural Resource Management, University of Copenhagen, Frederiksberg C, Denmark Katja Hansen EPEA Internationale Umweltforschung GmbH, Hamburg, Germany Department of Civil, Geo and Environmental Engineering, Technical University of Munich, Munich, Germany Volker Hartkopf Center for Building Performance and Diagnostics, Carnegie Mellon University, Pittsburgh, PA, USA Robert Hastings Danube University-Krems, Wallisellen, Switzerland Kevin Hydes Integral Group, Oakland, CA, USA Marina Bergen Jensen Landscape Architecture and Planning, Department of Geosciences and Natural Resource Management, University of Copenhagen, Frederiksberg C, Denmark Katrin Klingenberg Passive House Institute US, Chicago, IL, USA Manfred Köhler University of Applied Sciences, Neubrandenburg, Germany Helmut Köster Köster Lichtplanung, Frankfurt am Main, Germany Khee Poh Lam Center for Building Performance and Diagnostics, School of Architecture, Carnegie Mellon University, Pittsburgh, PA, USA School of Design and Environment, National University of Singapore, Singapore, Singapore Mary Ann Lazarus MALeco, Washington University, St. Louis, MO, USA Laura Lesniewski AIA, BNIM, Kansas City, MO, USA Vivian Loftness Center for Building Performance and Diagnostics, Carnegie Mellon University, Pittsburgh, PA, USA Pamela Mang Regenesis Group, Santa Fe, NM, USA Juliane Mathey Leibniz Institute of Ecological Urban and Regional Development (IOER), Dresden, Germany Douglas Mulhall EPEA Internationale Umweltforschung GmbH, Hamburg, Germany Department of Civil, Geo and Environmental Engineering, Technical University of Munich, Munich, Germany Department of Architectural Engineering + Technology, Delft University of Technology, Delft, The Netherlands Stephan Pauleit Center of Life and Food Sciences Weihenstephan, Technical University of Munich, Freising, Germany

Contributors

Contributors

xxi

Bill Reed Regenesis Group, Arlington, MA, USA Dieter Rink Helmholtz Centre for Environmental Research (UFZ), Leipzig, Germany Catherine O. Ryan Terrapin Bright Green, New York, NY, USA Robbert P. H. Snep Alterra, Wageningen University and Research Center, Wageningen, The Netherlands Megan Snyder School of Architecture and Center for Building Performance and Diagnostics, Carnegie Mellon University, Pittsburgh, PA, USA Donald Watson FAIA, Trumbull, CT, USA Alex Wilson Resilient Design Institute, Brattleboro, VT, USA Tong Yang Department of Design Engineering and Mathematics, Faculty of Science and Technology, Middlesex University, London, UK

Sustainable Built Environments: Introduction Vivian Loftness Center for Building Performance and Diagnostics, Carnegie Mellon University, Pittsburgh, PA, USA

Article Outline The Scale of the Sustainability Problem and the Opportunity Principles for the Design of a Sustainable Built Environment How Does This Broad Definition Translate into an Encyclopedia for Sustainable Built Environments? Models of Sustainable Built Environments (5 Chapters) Urban Design for Sustainability (6 Chapters) Building Design for Sustainability (9 Chapters) Sustainability Practices (5 Chapters) Conclusion for Sustainable Built Environments: Policy Matters – The Market Will Not Take Care of It Renewable Energy Generation and Sustainable Buildings References

For the moment, this is the definitive encyclopedia on sustainability in the built environment written by world leaders in each subject area. These authors have written the textbooks, the codes and standards, and the vision for a more sustainable built environment, at every level – regional, urban, community, building, and material. They have transformed practice, research, and education in fundamental ways, and it has been a privilege to work with them to assemble these 25 chapters.

The Scale of the Sustainability Problem and the Opportunity The building sector is the biggest “player” in the energy use equation and can have the greatest impact on maximizing energy supply and minimizing energy demand while providing measurable gains for productivity, health, and environmental quality. In the USA, commercial and residential buildings use 70% of total US electricity and are responsible for over 38% of total US greenhouse gas emissions [1] (Fig. 1, CBPD). It is critical for us to move away from diagrams that imply electricity is an end use sector, diagrams which overwhelm online searches. For prioritized investments in demand reduction, the electricity “sector” must be distributed to the true end use sectors – buildings, industry, transportation, and agriculture (Fig. 2a). The subdivision of the building sector alone into subsectors (residential and commercial) further masks the importance of this sector for policy, research, and investment as we strive to curtail global climate change (Fig. 2b). Equally significantly, however, are other environmental statistics related to the built environment: material use, water use, waste production, manufacturing and transportation costs, and carbon impacts. Buildings use 40% of the raw materials globally and 14% of the potable water in the USA. Building activity in the USA also contributes over 136 million tons of construction and demolition waste (2.8 lbs/person/day) and 38% of US greenhouse gas emissions [34]. These statistics underscore the impact of buildings and the built environment on global energy use and climate change, on water and material use. Through land use policies and actions, the built environment generated most of the increases in automobile transportation energy use in the last decades. The impact is even more substantial when human and environmental health, land use and community, as well

© Springer Science+Business Media, LLC, part of Springer Nature 2020 V. Loftness (ed.), Sustainable Built Environments, https://doi.org/10.1007/978-1-0716-0684-1_925 Originally published in R. A. Meyers (ed.), Encyclopedia of Sustainability Science and Technology, © Springer Science+Business Media, LLC, part of Springer Nature 2020, https://doi.org/10.1007/978-1-4939-2493-6_925-3

1

2

Sustainable Built Environments: Introduction

Sustainable Built Environments: Introduction, Fig. 1 The building sector is the greatest contributor to climate change, exceeding both industry and transportation

Sustainable Built Environments: Introduction, Fig. 2 CO2 and GHG charts must be corrected to clearly reveal that 39% of the environmental challenge is in

building construction and operations (replacing 2a with 2b and then 2c) [1, 2]

as indoor environmental quality are factored into sustainability. In short, the scale of the sustainability challenge in the built environment must include:

• The role of the built environment on material depletion • The role of the built environment on mobility and sustainable infrastructures

• The role of the built environment on global energy use and climate change • The role of the built environment on health and indoor environmental quality • The role of the built environment on water use

Principles for the Design of a Sustainable Built Environment Sustainable design is a collective process whereby the built environment achieves unprecedented levels

Sustainable Built Environments: Introduction

3

of ecological regeneration through new and retrofit construction, toward the long-term viability and humanization of architecture. Focusing on environmental context, sustainable design merges the natural, minimum resource conditioning solutions of the past (daylight, solar heat, and natural ventilation) with the innovative technologies of the present, into an integrated “intelligent” system that supports individual control with expert negotiation for resource consciousness. Sustainable design rediscovers the social, environmental, and technical values of pedestrian, mixed-use communities, fully utilizing existing infrastructures, establishing “eco-districts,” and recapturing indoor-outdoor relationships. Sustainable design avoids the further thinning out of land use and the dislocated placement of buildings and functions caused by single use zoning. Sustainable design introduces materials and assemblies fully integrated into a circular economy of materials as biological or technical nutrients. Finally, sustainable design offers architecture of long-term value through “forgiving” and modifiable building systems, through life cycle instead of least-cost investments, and through timeless delight and craftsmanship. A sustainable built environment depends on a number of critical principles:

to engineering to interior design to material science to social science and beyond. Sustainable built environments is a revolution for educational pedagogy and content. Moreover, it requires a level of trans-disciplinarity that challenges professional roles and creates new collaborative processes. To profile the breadth of science and engineering knowledge that is central to advancing sustainable built environments, multiple encyclopedia volumes would be needed to cover the breadth of subjects: public policy and economics of a sustainable built environment; architecture, mechanical engineering, lighting engineering, material engineering, water engineering, and energy engineering; as well as urban design, landscape, and urban infrastructure engineering. By necessity a single volume on sustainable built environments must offer a sampling of the science and technology expertise required. This encyclopedia has gathered 25 leading authors from around the world to provide insight into the breadth of knowledge needed to fully define and understand sustainable built environments. The 25 chapters are grouped in 4 major sections and ordered to build expertise across disciplines and scales of practice.

1. An integrative, human-ecological design approach 2. Changing approaches to land use and community fabric 3. The effective use of natural, local, and global resources to reduce infrastructure loading and maximize infrastructure use 4. The design of flexible systems, integrated for comprehensive performance delivery 5. The use of regenerative materials and assemblies 6. The design for life cycle instead of first cost 7. The promotion of infrastructures for water, energy, transportation, and connectivity to sustainable neighborhood amenities

• • • •

How Does This Broad Definition Translate into an Encyclopedia for Sustainable Built Environments?

Architectural and urban theory critically needs to embrace models of sustainable built environments as central to our professional and artistic explorations and to our collective future. The first set of articles in this encyclopedia introduce five of the enduring and emerging models for sustainable

A sustainable built environment spans dozens of professions, from urban planning to architecture

Models of sustainable built environments Urban design for sustainability Building design for sustainability Integrated delivery processes for sustainability

The following sections introduce these chapters within the broader knowledge base needed for substantially advancing our built environment from one that consumes and destroys to one that generates and revitalizes.

Models of Sustainable Built Environments (5 Chapters)

4

built environments: bioclimatic, biophilic, healthy, regenerative, and resilient. These leading authors helped to define the fields and have written and updated seminal encyclopedic entries. We must begin by recognizing that entire civilizations have successfully prospered in highly diverse and challenging climates without depleting natural resources beyond the ability of those resources to naturally regenerate. In large part, this success has been achieved through highly responsive, even intelligent building enclosures. Each climate commands a different strategy for natural conditioning, as so eloquently introduced in Don Watson’s entry on ▶ “Bioclimatic Design.” Emeritus Professor of Yale University and former Dean of RPI’s School of Architecture, Don Watson is a preeminent practitioner and scholar on the subject of Climate and Architecture and author of standards and textbooks used worldwide, beginning with his McGraw-Hill textbook Climatic Building Design: Energy Efficient Building Principles and Practices [3] through his most recent John Wiley textbook, Design for Flooding: Architecture, Landscape, and Urban Design for Resilience to Climate Change [4]. Bioclimatic design embraces the imperative that nature can and must be the dominant conditioning system, with mechanical conditioning as a welldesigned partner, in use only when needed. The benefits of bioclimatic design go well beyond energy savings to include environmental and sensory richness in the spaces we occupy, climatic and cultural diversity in the built environment, and significantly enriched social lives. A companion to bioclimatic design is ▶ “Biophilic Design,” in which nature is a dominant sensory system for enriching the quality of life. The authors, Catie Ryan and Bill Browning of Terrapin Bright Green, are world leaders in the theory, science, and practice of biophilic design, leveraging expertise in high-performance design, whole-systems thinking, bioinspired innovation, and ecology [5]. Building on the foundational work of E.O. Wilson and Stephen Kellert [6], this chapter introduces Terrapin’s 14 patterns of biophilic design with rich illustrations alongside worldwide research on the benefits of biophilia for health and well-being (https://www.terrapin brightgreen.com/report/14-patterns/). Biophilic

Sustainable Built Environments: Introduction

design, in combination with bioclimatic design, is essential to reconnecting people with their unique environments leading to a healthy, prosperous, and regenerative future for all. Both bioclimatic and biophilic design are driven to go beyond sustainability to enhance quality of life, health, and well-being. The importance of physiological, psychological, and social health as a model for design merits a chapter of its own. In ▶ “Sustainable and Healthy Built Environment,” Vivian Loftness and Megan Snyder introduce a framework for ensuring that building and communities are designed to achieve the highest level of human health outcomes while generating the lowest environmental footprint [7]. Across 30 years of research and teaching in the School of Architecture at Carnegie Mellon, the authors have been developing a database of research linking building components and systems, as well as land use approaches and infrastructures, to health, productivity, quality of life, as well as other economic outcomes of critical importance to decision-makers. While bioclimatic, biophilic, and healthy models for sustainability focus on the human impacts of the built environment, regenerative design focuses on fully addressing the environmental impacts as well. Sustainable design models continue to mature, from energy and water, to indoor environmental quality and equity, to models that embrace living buildings, restorative buildings, and regenerative buildings. Two international leaders, Pamela Mang and Bill Reed, outline the regenerative model for design in the chapter ▶ “Regenerative Development and Design.” Building on the early principles of permaculture and the regenerative design principles of John Lyle [8] and David Orr [9], they have advanced regenerative design through their integrated practice Regenesis, Inc. Co-author with the Seven Group of The Integrative Design Guide to Green Building published in 2009 [10] by John Wiley, Bill Reed and Pamela Mang introduce the integrative, wholesystems regenerative design process to ensure that “projects and processes achieve the highest level of performance across every sustainability metric – and beyond, into regenerating the health of our ecological systems [11].”

Sustainable Built Environments: Introduction

The outward spiraling models of sustainability continue to evolve in the face of climate change, leading to the fifth model of vital importance to sustainable built environments – ▶ “Resilient Design” – written by Alex Wilson and MaryAnne Lazarus. Founded at BuildingGreen under Alex Wilson’s leadership, the non-profit Resilient Design Institute is dedicated to providing expertise for “the intentional design of buildings, landscapes, communities, and regions in order to respond to natural and manmade disasters and disturbances – as well as long-term changes resulting from climate change – including sea level rise, increased frequency of heat waves, and regional drought (https://www.resilientdesign.org/).” These five models for sustainability – bioclimatic, biophilic, healthy, regenerative, and resilient – as well the “Cradle to Cradle” model captured in the final chapter of this encyclopedia [25], are essential to design education, design practice, public policy, and public and private investment. They are also critical to setting the stage for standards, research, and innovation to ensure that the built environment is no longer the cause of depleting resources and reduced health, but a source of design responses for ensuring restorative human and ecological futures.

Urban Design for Sustainability (6 Chapters) Most sustainable designers have long since realized that if the urban design decisions are not sustainable, the buildings themselves will not make the decisive difference for human and ecological futures. The design of land use and landscapes for regenerative sites, neighborhoods, and entire regions has the potential to address a broad range of environmental challenges: water shortages, flooding, brownfields, carbon sequestration, habitats, access to food, and human health. Through collaborative design, the building community should fully engage the science and technology of: • Land use planning for farm and forest preservation

5

• Land use planning for walkability and quality of life • Watershed planning • Landscape design for net-zero water and aquifer recharge • Cool roofs and cool communities • Landscape design for brownfield remediation • Landscape design for climate mitigation and carbon sequestration • Building and landscape design for food, biotopes, and habitats The six chapters in this section introduce critical shifts needed in urban design to ensure a regenerative future that addresses the challenges of energy, water, mobility, food, community, and more. The definitions must begin with land use planning. Douglas Farr of Farr Associates in Chicago is the author of the book on Sustainable Urbanism [12], evolving from his years of leadership as chair of LEED ND and the board of the Congress for New Urbanism. In the chapter ▶ “Sustainable Urbanism,” he outlines key design actions for sustainable land use and community design. He contends that sustainable urbanism must embrace walkable and transit-oriented living, integrated with high-performance buildings and infrastructures. Compactness (density), completeness, connectedness, and biophilia integrate human and natural systems to dramatically improve energy, water, waste, food, and mobility for a sustainable future. The architectural and urban design imperatives for a sustainable future must be matched by the landscape and engineering imperatives for the underlying resources and infrastructures that support sustainable urbanism. The most demanding set of design imperatives center on water. Global demands for water are depleting freshwater availability at a rate that is alarming. While 70% of global freshwater demand is for agriculture, the 20% freshwater used in the building sector is growing at a rapid rate – with the highest level of inefficiency. As with energy, the greatest opportunities for conservation at the lowest first cost and the highest payback reside in the building sector. UNEP illustrates this with charts that differentiate

6

Sustainable Built Environments: Introduction

water extraction and consumption per end use sector (see Fig. 3), highlighting the major opportunities for the building sector to dramatically reduce its extraction demands through efficiency, gray water, and black water innovations [13]. Herbert Dreiseitl, of Atelier Dreiseitl and Ramboll, is the world’s leading water designer, combining the skills of an artist with the expertise of a landscape architect and a water infrastructure engineer in innovative projects around the world. As the author of Recent Waterscapes [14] and Waterscape Innovations [15], his chapter ▶ “Water and Sustainable Design” introduces the importance of comprehensively understanding

watershed, the natural path of rainfall and stream flows given the topographic and geologic construct of the land, flowing into rivers, lakes, wetlands, or the sea. To address the challenges of deluge, drought, and contamination, the chapter illustrates the power of green infrastructures and blue-green infrastructures as an approach to water management that protects, restores, and emulates the natural water cycle. Combining natural and technical water systems that ensure the creation of water-sensitive, regenerative cities, the chapter introduces critical theory and practices, richly illustrated by projects in Asia, Europe, and the USA.

Sustainable Built Environments: Introduction, Fig. 3 While agriculture represents the largest consumer of freshwater worldwide, the greatest waste in freshwater

use is the built environment, shown in the difference between extraction and consumption

Sustainable Built Environments: Introduction

Green-blue infrastructures have an equally important role to play in climate mitigation and adaptation. In their chapter ▶ “Green Infrastructures to Face Climate Change in an Urbanizing World,” Stephan Pauleit of the Center of Life and Food Sciences at the Technical University of Munich, and Ole Fryd, Antje Backhaus, and Marina Bergen Jensen of the University of Copenhagen quantify the role of green infrastructures in cities for climate adaptation and the potential for climate mitigation, given the vulnerability of cities and their infrastructures to climate change. They outline the combined challenges of hazard, vulnerability, and exposure in relation to urban heat, storm water management, flood management, and carbon sequestration. The solution sets are rich assemblies of urban morphology solutions that mitigate climate change through annual carbon sequestration rates as high as 130 kgCO2/ha, as well as adapt to climate change through low-impact design, water-sensitive urban design, and storm water management. The challenge of shrinking cities in both Europe and the USA, in which abandoned properties and vacant sites leave a “perforated” urban fabric, is a major opportunity to create permanent green spaces and green corridors, reintroducing biodiversity, farm, and forest into the city. The opportunity and approaches to re-naturalizing vacant brownfields as green infrastructure – for ecological upgrading, nature and species conservation, climate change mitigation, and quality of life improvements – are introduced in the chapter ▶ “Greening Brownfields in Urban Redevelopment” by Juliane Mathey of the Leibniz Institute of Ecological and Regional Development in Dresden and Dieter Rink of the Helmholtz Centre for Environmental Research in Leipzig. In addition to the benefit of climate change mitigation and quality of life improvements, blue-green infrastructures play a central role in regenerating biodiversity in cities. Worldwide, the diversity of plant and animal life is diminishing rapidly, with the latest global Living Planet Index of the World Wildlife Foundation showing a decline of biological diversity of 60% between 1970 and 2014 (http://www.livingpla netindex.org/home/index).

7

In the chapter ▶ “Biodiversity in Cities, Reconnecting Humans with Nature,” Robert Snep of Wageningen University in the Netherlands and Pierre Clergeau of the Department of Ecology and Biodiversity of the Museum of Natural History in Paris identify the wide range of services provided by urban biodiversity that are of vital importance for the quality of human life. They propose a commitment to habitat corridors and green space connectivity to ensure city environments that support wildlife and plant diversity, integrating wildlife and biodiversity regeneration in sustainable city planning and design. They define cities as a “Mosaic of Biotopes” – vital to species diversity and survival. The final chapter in this set is the seminal work on ▶ “Biophilic Cities” by Timothy Beatley, the Teresa Heinz Professor of Sustainable Communities in the Department of Urban and Environmental Planning at the University of Virginia. Author of 15 books, his 2010 book Biophilic Cities: Integrating Nature Into Urban Design and Planning [16] and his 2017 Handbook of Biophilic City Planning & Design [17] have been instrumental in the global movement to bring the essential elements of biophilic abundance to every city. This chapter defines the inspiration and criticality of biophilia in urban design for sustainability: cities of abundant nature conserving and protecting biodiversity; cities embedded in nature and natural systems; cities that understand the healing powers of nature; and cities of wonder, awe, and fascination.

Building Design for Sustainability (9 Chapters) Building design, renovation, and management is the next level of critical action for human and ecological restoration. The next nine chapters provide a compendium of expertise on the sweep of design choices, from adaptive building enclosures to adaptive mechanical systems, to embracing daylight, natural ventilation, and sunshine, to rethinking our roofs and foundations. In each case, world leaders in research and application of

8

these disciplines have authored an invaluable encyclopedic entry. In ▶ “Facades and Enclosures: Building for Sustainability,” Dr. Volker Hartkopf and Senior Researcher Azizan Aziz build on each climate’s imperatives to introduce the richness of design responses for sustainable building design. Professor of Architecture and Director of the Center for Building Performance and Diagnostics at Carnegie Mellon University, Volker Hartkopf has led an industrial and government consortium dedicated to advanced building systems integration for performance through collaborative research, demonstration, and policy for over 30 years. This leadership led to the construction of the Intelligent Workplace, a living laboratory for innovation in building materials, components, and integrated systems for improving human comfort and health, individual and organizational productivity, and environmental sustainability [19]. This chapter defines the enclosure science and design innovations that are pivotal to ensure the highest performance outcomes for each climate and building type. With over 12 distinct fields of design from plenum to clerestory and interior to exterior layers, high-performance fac¸ades must balance access to the natural environment for daylighting, natural ventilation, thermal mass, and passive solar heat, with heat loss/heat gain control, solar control and enclosure advances for water management, material integrity, and advanced systems integration. For most of the northern climates, heating is the largest energy demand and carbon footprint for buildings, arguing for design and retrofit innovations that combine deep energy conservation with the free energies of sunshine. The importance of passive solar design began in the 1970s after the first energy crisis, giving rise to today’s Passivhaus standards emerging throughout Europe, Scandinavia, and the USA. The chapter on ▶ “Passive House (Passivhaus)” is authored by Katrin Klingenberg, the Co-founder and Director of the Ecological Construction Laboratory and the Passive House Institute in the USA (PHIUS). A long-term collaborator with the Passivhaus Institute in Europe, she co-authored the book Homes for a Changing Climate: Passive Houses

Sustainable Built Environments: Introduction

in the US [19], to emphasize both the potential of passive house design to dramatically reduce or eliminate thermal conditioning in homes and the importance of climate-specific approaches to balancing conservation and passive conditioning. While based on deep energy conservation through insulation and air tightness, the power of solar energy for low-energy heating must not be understated. The trajectory of the passive solar design movement is richly illustrated in Professor Robert Hastings ▶ “Passive Solar Heating in the Built Environment.” Emeritus Professor of the École Polytechnique Fédérale de Lausanne and Managing Director of Architecture, Energy and Environment GmbH in Switzerland, Robert Hastings is a leader in the IEA Task 28-Solar Heating and Cooling. A prolific author with a long list of technical publications on energy efficiency, he co-edited Solar Energy Houses: Strategies, Technologies and Examples published by Earthscan [20]. This chapter is key to understanding the systemic nature and diversity of passive solar heating in use throughout Europe and North America. Lighting is the second largest energy demand in buildings. Over 10% of all US energy is used for lighting buildings, much of this during the daytime when daylight is abundant. Effective daylighting can yield 30–60% reductions in annual lighting energy consumption. Helmut Koester’s entry on ▶ “Daylighting Controls, Performance, and Global Impacts” addresses the design and engineering innovations needed for ensuring daylight’s massive contribution to energy and carbon savings. The author of Dynamic Daylighting Architecture: Basics, Systems, Projects published by Birkhäuser Press [21], Helmut Koester introduces the science and beauty of external and internal fac¸ade layers for effective daylighting and view, as well as shading, glare, and brightness contrast control. If complemented by cuttingedge electric lighting and lighting control systems, the residual energy demands for lighting would be less than 10% of today’s demand, an amazing 6–8% reduction in our global carbon footprint. The rapid growth in building air conditioning in the face of global warming makes natural

Sustainable Built Environments: Introduction

cooling and natural ventilation equally important for a no or low carbon built environment. In ▶ “Natural Ventilation in Built Environment,” Tong Yang and Derek Clements-Croome illustrate the potential for natural ventilation and natural cooling to ensure the highest level of thermal comfort and air quality at the lowest energy cost – through design and engineering innovation. Editor of Intelligent Buildings International Journal and Emeritus Professor of the University of Reading, Derek Clements-Croome is the CIB Coordinator for WO98 Intelligent and Responsive Buildings and editor of Naturally Ventilated Buildings: Buildings for the Senses, the Economy and Society, published by E & F N Spon [22]. While deep conservation and passive heating, cooling, and daylighting are the critical first steps, innovation in building mechanical systems for sustainability is equally important. Kevin Hydes, CEO of the Integral Group, and Jennifer Fosket of Social Green introduce the breadth and importance of sustainable mechanical engineering, and the changes in practice that will be vital to a sustainable future, in the entry ▶ “Sustainable Heating Ventilation and Air Conditioning.” They argue that “the best mechanical system is one that you never need to turn on” calling for dramatic changes in the role of the mechanical engineers in practice. A founder of the Canadian Green Building Council, Past Chair of the Board of Directors of the USGBC and the World Green Building Council, Kevin Hydes is central to the transformative reference Ecological Engineer from Ecotone Publishing [23]. In this chapter, they demonstrate how building mechanical systems can be systematically rethought for sustainability, embracing waterbased thermal conditioning, dedicated outside air and displacement ventilation, energy recovery, innovations in building automation, integrated passive conditioning and renewable energy, and commissioning. Integral to advanced HVAC design, innovative architects and engineers are shifting to a hybrid approach to space conditioning that uses natural ventilation from operable windows (either manually or automatically controlled) or other passive inlet vents, in combination with mechanical systems to provide the critical mix of air distribution

9

and cooling. Richard de Dear of the University of Sydney and Gail Brager of the University of California at Berkeley have published widely on adaptive comfort and led the movement to bring adaptive comfort and mixed-mode conditioning into mainstream codes and standards (ASHRAE Standard for Natural Ventilation [24]). Their chapter ▶ “Adaptive Comfort and Mixed-Mode Conditioning” introduces the adaptive thermal comfort models and proof sets, their integration into mixed-mode buildings, and their implications for environmental quality and energy use. A welldesigned mixed-mode building allows spaces to be naturally ventilated during periods of the day or year when it is feasible or desirable, and uses air conditioning for supplemental cooling when natural ventilation is not sufficient through seasonal changeover, concurrent, or zoned modes of operation (https://escholarship.org/uc/item/3f73w323, https://cbe.berkeley.edu/research/adaptive-comfortmodel/). Mixed-mode conditioning enhances thermal comfort and health while minimizing the significant energy use, carbon footprint, and operating costs of air conditioning. Two additional perspectives are offered in this set of entries on building systems for sustainability – focused on the fifth fac¸ade, roofs, and the sixth, foundations. Manfred Koehler of the University of Neubrandenburg and Andrew Michael Clements of Green Roof Greece introduce the imperative for rapid increases in extensive and intensive green roofs as urban ecosystems in the entry ▶ “Green Roofs: Ecological Functions of the Fifth Facade.” They outline the main ecological and environmental challenges to be met by green roofs – addressing water cycles and urban water management, urban heat and energy demands, and losses in urban biodiversity – as well as quantifying the measurable ecological and economic benefits that are derived by widespread greening of urban roofs. While other volumes of the Springer Encyclopedia offer definitive chapters on combined heat and power, photovoltaics, and solar thermal systems at both the building and community level, there is one renewable energy system that is fully integrated into the building design and engineering – geothermal heating and cooling. Co-author

10

of a definitive text from the National Ground Water Association, Carnegie Mellon Professor Nina Baird brings the most cutting-edge knowledge of the breadth and depth of the engineering challenges and opportunities on open- and closedloop geothermal energy systems in the entry ▶ “Geothermal Conditioning: Critical Sources for Sustainability.” From earth sheltering to earth tubes, from horizontal and vertical borehole heat exchangers, to aquifer and wastewater-based geothermal systems, ground-coupled heating and cooling systems are invaluable low energy and low carbon solutions to conditioning individual buildings, campuses, and entire communities.

Sustainability Practices (5 Chapters) The final five chapters define the changes in delivery processes needed to ensure the integrated design and engineering innovations for sustainability. These authors contend that urban and building sustainability will be critically dependent on new forms of professional practice that: • Engage the entire building delivery team at the outset • Set sustainability goals and meet global standards of excellence • Effectively use performance simulation tools throughout design, delivery, and operation • Engage in post occupancy evaluation with qualitative and quantitative feedback and continuous improvement • Pursue closed-loop, Cradle to Cradle™ design While each of these chapters defines significant changes for delivering buildings and communities, they also represent theoretical models for advancing sustainable practices, with as much urgency as the six that began this volume of the encyclopedia. In ▶ “Sustainable Design and Construction, Integrated Delivery Processes, and Building Information Modeling,” Laura Lesniewski, a principal in the leading sustainability firm of BNIM, and Bob Berkebile a founder of this firm and

Sustainable Built Environments: Introduction

many sustainability standards in the USA, redefine integrated practices for sustainability. With a significant portfolio of LEED, net-zero, and Living Building Challenge projects, they bring important insight to the changes needed in professional practice for sustainability, from early multidisciplinary collaboration to the most advanced building information modeling to embrace the science and technology of sustainability. They introduce sustainable design projects as a network of commitments with overlapping and iterative improvements to optimize the entire project instead of subsystems. Raymond Cole, Professor at the University of British Columbia and Head of the Environmental Research Group (ERG), provides a worldwide overview of ▶ “Rating Systems for Sustainability.” The chapter introduces sustainability assessment methods at multiple scales, alongside voluntary and regulatory standards and practices from around the world. One of the leading experts on sustainable building practices, Dr. Cole has helped to forge the evolution in standards to ensure broad environmental responsibilities for designers – from energy to water to materials to indoor environmental quality to land use – a breadth that has been critical to defining sustainability science and technology for the built environment. Dr. Khee Poh Lam, Dean of the School of Design and Environment at the National University of Singapore (NUS), and Emeritus Professor at Carnegie Mellon University, has written the decisive entry on ▶ “Sustainability Performance Simulation Tools for Building Design.” His leadership in the development and application of energy, lighting, and computational fluid dynamics tools has refined building and urban projects around the world, including a new Net-Zero campus building at NUS. His development of global standards for interoperability, and design compliance tools for sustainable practices, contributes to this expert overview of computational tools and their contributions to ensuring measured environmental outcomes. Sustainability in the built environment is equally dependent on field data collection through post occupancy evaluation (POE) for iterative and

Sustainable Built Environments: Introduction

continuous improvements in design and operation. Dr. Charlene Bayer provides an introduction to these feedback loops in the chapter ▶ “Evidence-Based Design for Indoor Environmental Quality and Health.” As Leader of the Environmental Exposures and Analysis Group at the Georgia Tech Research Institute (GTRI) and Director of Georgia Institute of Technology’s Indoor Environment Research Program, Dr. Bayer brings over 30 years of experience to definitions of indoor environmental quality. She introduces the power of a field research-based approach to facilities design “that treats the building and its occupants as a system and gives importance to design features that impact health, well-being, mood and stress, safety, operational efficiency, and economics.” The contributions of Evidence-based Design (EBD) to ensuring sustainable built environments are based on shifting the focus from the designed performance to the installed performance of integrated building systems and re-centering goals to focus on the productivity and health of building occupants. There is not a more significant closure to these 25 chapters than ▶ “Materials Banking and Resource Repletion, Role of Buildings, and Materials Passports” by Katja Hansen, Michael Braungart, and Douglas Mulhall. This chapter fully bridges the four thematic areas by introducing the critical model for embodied energy and materials that is redefining sustainable built environments yet again. As the sustainable design community moves toward carbon neutral buildings and communities, the impact of building materials and their embodied energy have never been more important. The manifesto Cradle to Cradle [25] and more recently Upcycle [26] by William McDonough and Michael Braungart are two of the essential references for our time, challenging professionals to treat all materials as industrial or agricultural nutrients in closed cycles. Braungart, Hansen, and Mulhall bring the full expertise of three university disciplines, EPEA International Umweltforschung GmbH, and McDonough Braungart Design Chemistry (MBDC) to address the embodied carbon, health, and raw materials used in an “economy that will run out of materials before it runs out of energy.”

11

The authors begin with the goals that all materials are designed to be nutrients in a circular economy, yet expand the model to introduce material passports and nutrient certificates as essential to a sustainable and regenerative future.

Conclusion for Sustainable Built Environments: Policy Matters – The Market Will Not Take Care of It Energy, water, and materials are too cheap, privatizing the profits while socializing the “externalities” of resource depletion, climate change, pollution, and quality of life. In the USA, governments, building owners, and consumers do not see energy as a large enough component of their disposable income to even evaluate the return on investing in energy efficiency, much less water efficiency, or other sustainability goals. Deregulation has reduced the efforts of major utilities to pursue demand side management and weatherization. At the same time, power unreliability has led residential and commercial building owners to purchase inefficient and polluting standby power rather than consider the significant opportunity to invest in energy efficiency and renewable energy sources. Policy is critical because the market will not act quickly or forcefully. The contributions of buildings to the discharge of four primary pollutants – NOX, SOX, CO2, and particulates – should be fully recognized in the cost of energy, to catalyze owners and occupants to pursue more environmentally responsible buildings and building use patterns. Federal and state energy efficiency standards, as well as tax incentives, are key to finally balancing investments between energy supply and energy demand. Given the wasteful excesses in energy use in the built environment, reducing demand must instead be seen as a major energy source. The McKinsey report “Unlocking Energy Efficiency in the US Economy” was a revelation for the energy world, finding that carbon reductions through building energy efficiency are the lowest cost per ton with the highest return on investment of any greenhouse gas (GHG) action (Fig. 4, [27]). Investments in “mining” this new energy supply will yield greater economic benefit for a broader

12

Sustainable Built Environments: Introduction

Sustainable Built Environments: Introduction, Fig. 4 Carbon reductions through building energy efficiency are the lowest cost per ton with the highest return on investment (http://www.mckinsey.com)

array of industries; provide significant gains in reducing environmental pollution; and ensure a longevity to energy “supply” that few other sources can ensure. Even though the return on investment for energy efficiency dramatically exceeds that of creating new sources, the US dollars in energy supply research and development (R&D) are seven times the R&D dollars for research in energy demand reductions (https://www.energy.gov/sites/prod/files/2019/ 03/f60/doe-fy2020-budget-in-brief_0.pdf). Yet the modest national investments by the US DOE in R&D for energy-efficient ballasts, low-E windows, and refrigerator standards reaped national benefits of $9,000, $7,000, and $23,000 per $1 dollar invested [28]. In addition to the obvious benefits of reduced energy demand, dramatically accelerated national investments and policies focused on building energy efficiency will contribute to the following: • Reduced unnecessary consumption

annual

energy

• Reduced emissions and climate change impacts • Increased peak power capacitance and reliability (energy security) • Improved health, human safety, and security • Improved productivity • Improved quality of life • Increased exports of products and services • Setting a proven example for emerging nations with growing demands Setting a proven example in mitigating climate change through sustainable building and land use is especially important for emerging economies that often assume the need for expanding energy demand alongside development. Greg Kats argues in a study of the costs and financial benefits of green buildings, “The vast majority of the world’s climate change scientists have concluded that anthropogenic emissions - principally from burning fossil fuels - are the root cause of global warming. . . The US building sector is responsible for about 35% of US CO2 emissions, the dominant

Sustainable Built Environments: Introduction

global warming gas” [29]. The actions needed include massively improving building energy efficiency, building land use, and infrastructure efficiency – in the very near term. This can only be achieved by immediate changes in policy, investment, and research at the national, regional, and industrial level. There is ample demonstration that energy efficiency in buildings represents a major untapped resource for energy supply. The development of national standards and the removal of market barriers can lead to significant reductions in energy use from key building technologies through their natural replacement cycle. Despite massive increases in residential refrigeration, for example, standards have effectively reduced energy demand to 1947 levels ([30, 31], Fig. 5). The lack of sustained investment in energy demand research and development makes a 1997 study, undertaken by all five national laboratories quantifying the potential, still actionable 20 years later. They determined that building energy efficiency could achieve 60% of the energy savings needed to meet US targets under the Kyoto

13

Protocol. With the addition of innovative combined cooling, heat, and power technologies, a further 170MTC could be achieved, meeting 100% of the 2010 goals through the building sector alone [32]. Investing in building energy efficiency as a new energy “supply” would dramatically surpass production from new oil supplies and power plant investments, as well as offer sustained “sources” of energy that do not generate greenhouse gases. Yet the combined budgets for building research across US government agencies is less than 2% of federally funded R&D, in no way commensurate with the importance of the built environment to our economy and quality of life [33]. Given this paucity of research support, there are only a handful of university Ph.D. programs in the USA focused on energy efficiency and environmental quality in the built environment, an inadequacy that leaves the US industry lagging in its ability to innovate for building energy efficiency. Over the longer term, expanded building R&D budgets, industry- and university-based research, and continuing national policies that focus on

Sustainable Built Environments: Introduction, Fig. 5 Driven by California, national standards led to significant reductions in energy use from key buildings technologies – a major “source” of new energy for the USA

14

building energy efficiency could trigger dramatic improvements in energy and environmental quality in the built environment. Moreover, these investments would ensure ancillary benefits including the revitalization of existing buildings and infrastructures, measurable gains in health and productivity, and a positive influence on energy-efficient growth in the built environment of developing nations.

Renewable Energy Generation and Sustainable Buildings Innovation in sustainable buildings and community design and operation must also be integral to the supply side of the equation. In 2002, architect Edward Mazria issued the 2030 challenge [34], asking the global architecture and building community to adopt 60% energy savings goals by 2010, 70% by 2015, advancing each 5 years by 10% to achieve carbon neutrality by 2030 for all new buildings and major renovations (http:// www.architecture2030.org/). The 2030 challenge demonstrated that these targets can be accomplished by “implementing innovative sustainable design strategies, generating on-site renewable power, and purchasing renewable energy (up to a 20% maximum).” In the USA, over 500 architectural firms have committed to these goals for their entire portfolios, with projects representing a savings of 17.7mT of CO2e in 2018 alone (20.8 million acres of forest carbon sequestration). Today, the challenge has been amended to reduce embodied carbon emissions from all new buildings, infrastructure, and associated materials as well – 50% by 2030 to zero by 2050. While demand reduction is the critical first step, sustainable alternatives to energy supply must also be fully integrated into community and building design. While these are the subject of other volumes of this encyclopedia, buildings and communities play a significant role in decarbonizing energy supply through: • Electrification of the built environment after the lowest conditioning, process, plug, and parasitic loads are achieved through conservation and energy cascades;

Sustainable Built Environments: Introduction

• Combined heat and power (CHP) and building integrated CHP; • Site and building integrated photovoltaics (PV) • Site and building integrated solar thermal; • Geothermal, aqua-thermal, and groundcoupled HVAC. In pursuing these innovative energy sources, it is vital for the design community to aggressively pursue both “exergy” and energy cascades. In their PLEA 2004 paper, “Critical analysis of exergy efficiency definitions applicable to buildings and building services,” Boelman and Sakulpipatsin discuss the importance of matching each type of energy (electricity, low and high temperature thermal), and their energy potential, to their thermodynamically optimal uses in buildings and building services [35]. Exergy begs the question “if the Earth provides moderate temperature cooling and the Sun provides moderate temperature heating, can we design thermal conditioning systems that use these sources directly with maximum effect, leaving electricity for more challenging tasks?” At the same time, a sustainable built environment would use the building as a central element in energy cascades, using the “waste” heat from one process for the next demand and using the building mass as a “battery” for thermal energy. In conceiving the “Building as Power Plant,” Volker Hartkopf suggests that each new building should be a net energy exporter: first combining energy efficiency with passive conditioning in an “ascending” strategy of daylighting, natural ventilation, passive cooling, and passive solar heating and then integrating “cascading” energy uses, with the waste heat from power generation cascading to cooling energy, waste heat from cooling cascading to heating energy, and waste heat from heating cascading to hot water through the integrated design of the building, including the management of peak loads and time shifts for each energy source [36] (Fig. 6). Given that buildings in the USA account for 70% of US electricity consumption, 40% of total energy use, 38% of carbon dioxide (CO2) emissions, 40% of raw material use, 30% of waste output (136 million tons annually), and 14% of potable water consumption and are significantly

Sustainable Built Environments: Introduction

15

Sustainable Built Environments: Introduction, Fig. 6 Carnegie Mellon’s vision for a “Building as Power Plant” uses energy cascades to meet power, cooling, heating, and hot water demands, matching loads to optimize the “exergy” of each source, with excess energy for export to the campus grid [18, 36]

linked to health and competitiveness [37], it is vital to invest in policy, research, and innovation for more sustainable built environments. While Sustainable Cities and Communities is one of the key United Nations Sustainable Development Goals [38] (UNSDG 11 https://sustainable development.un.org/sdg11), this Springer volume illustrates that the built environment is equally pivotal for almost every goal: clean water; affordable and clean energy; decent work and economic development; industry, innovation, and infrastructure; reduced inequalities; responsible consumption and production; climate action; life on land; and partnerships to achieve these goals. Working together, all nations need to enact collaborative efforts to ensure that decision-making in the built environment contributes to a shared future of regenerative energy, air, water, land, and material resources while ensuring a sustainable quality of life for all.

3.

4.

5. 6.

7.

8.

9.

References 10. 1. U.S. EPA (2019) Draft inventory of U.S. Greenhouse Gas Emissions and Sinks 1990–2018, EPA 430-P-20001, and EIA Annual Energy Outlook 2020. https:// www.eia.gov/outlooks/aeo/ and Annual Energy Review http://38.96.246.204/totalenergy/data/annual/ index.cfm 2. Rock M, Saade M et al (2020) Embodied GHG emissions of buildings – the hidden challenge for effective

11.

12.

climate change mitigation. Appl Energy 258:114107. Elsevier Publishing☆ Watson D, Labs K (1993) Climatic building design: energy-efficient building principles and practices. McGraw Hill, New York. ISBN 10: 007068488X Watson D, Adams M (2012) Design for flooding: architecture, landscape, and urban design for resilience to flooding and climate change. Wiley, Hoboken. https://doi.org/10.1002/9781118259870 Browning B, Ryan C (2020) Nature inside: a biophilic design guide. RIBA, London ISBN-10: 1859469035 Kellert S, Heerwagen J, Mador M (2013) Biophilic design: the theory, science, and practice of bringing buildings to life. Wiley, Hoboken. ISBN: 978-0-47016334-4 Loftness V, Harktopf V (2002) Building investment decision support (BIDS): cost-benefit tool to promote high performance components, flexible infrastructures and systems integration for sustainable commercial buildings and productive organizations. The Austin Papers. Building Green, Brattleboro VT. ISBN 1-929884-10-9 Lyle J (1996) Regenerative design for sustainable development. Wiley, Hoboken. ISBN: 978-0-47117843-9 Orr D (2004) Earth in mind: on education, environment, and the human prospect. Island Press, Washington, DC. ISBN-10: 1559634952 7 group, Reed B (2009) The integrative design guide to green building: redefining the practice of sustainability. Wiley, Hoboken. ISBN-10: 0470181109 Mang P, Haggard B, Regenesis Group (2016) Regenerative development and design: a framework for evolving sustainability. Wiley, Hoboken. ISBN: 978-1-118-97286-1 Farr D (2007) Sustainable urbanism: urban design with nature. Wiley, Hoboken. ISBN 10: 047177751X

16 13. UNEP (2009) Vital water graphics: an overview of the state of the world’s fresh and marine waters. State Hydrological Institute/UNESCO, St. Petersburg/ Paris. http://wedocs.unep.org/bitstream/handle/20. 500.11822/20624/Vital_water_graphics.pdf 14. Dreiseitl H (2009) Recent waterscapes: planning, building and designing with water, 3rd edn. Birkhauser, Basel. ISBN-10: 3764389842 15. Dreiseitl H, Grau D (2014) Waterscape innovations. Design Media Publishing, Hong Kong. ISBN-10: 9881296935 16. Beatley T (2010) Biophilic cities: integrating nature into urban design and planning. Island Press, Washington, DC ISBN-10: 1597267155 17. Beatley T (2017) Handbook of biophilic city planning & design. Island Press, Washington, DC ISBN-10: 1610916204 18. Resnick D, Scott D (2004) Innovative university. CMU Press, Pittsburgh. ISBN 10: 0887483763 19. Klingenberg K, Kernagis M, James M (2009) Homes for a changing climate: passive houses in the U.S. Aspen, New York 20. Hestnes AG, Hastings R, Saxhof B (2003) Solar energy houses: strategies, technologies, examples. Earthscan, London. ISBN 10: 1902916433 21. Koster H (2004) Dynamic daylighting architecture. Birkhauser Architecture, Basel. ISBN: 9783764367305 22. Derek C (1998) Naturally ventilated buildings: building for the senses, the economy and society. Spon Press, New York. ISBN 10: 0419215204 23. Macauly D, McLennan J (2005) The ecological engineer. Ecotone Publishing, Kansas. ISBN 10: 0974903345 24. Brager G, de Dear R (2001) Climate, comfort and natural ventilation: a new adaptive comfort standard for ASHRAE standard 55. https://escholarship.org/uc/ item/2048t8nn 25. McDonough W, Braungart M (2002) Cradle to cradle: remaking the way we make things. North Point Press, New York. ISBN 10: 0865475873 26. McDonough W, Braungart M (2013) The upcycle: beyond sustainability – designing for abundance. North Point Press, New York. ISBN10: 0865477485 27. McKinsey Report (2013) Pathways to a low carbon economy: version 2 of the greenhouse gas abatement

Sustainable Built Environments: Introduction

28.

29.

30.

31.

32.

33.

34. 35.

36. 37.

38.

cost curve. https://www.mckinsey.com/~/media/ McKinsey/Business%20Functions/Sustainability/Our% 20Insights/Pathways%20to%20a%20low%20carbon% 20economy/Pathways%20to%20a%20low%20carbon %20economy.ashx Rosenfeld AH (1999) The art of energy efficiency: protecting the environment with better technology. Annu Rev Energy Environ 24:33–82 Kats G (2009) Greening our built world: costs, benefits, and strategies. Island Press, Washington, DC. ISBN 10: 159726668X Rosenfeld AH, McAuliffe P, Wilson J (2004) Energy efficiency and climate change. In: Braungart M (ed) Encyclopedia on energy. Cutler Cleveland, Academic, Elsevier Science Rosenfeld AH (2003) Improving energy efficiency 2–3% year to save money and avoid global warming. In: Sessler Symposium, LBNL, 15 March 2003, Berkeley Inter-laboratory Working Group (1997) Scenarios of US carbon reductions: potential impacts of energy technologies by 2010 and beyond. LBNL 40533. https://eta.lbl.gov/publications/scenarios-us-carbonreductions USGBC (2007) A national green building research agenda. http://www.usgbc.org/ShowFile.aspx? DocumentID¼3402. Accessed 20 Mar 2020 Mazria E (2002–present) The 2030 challenge. http:// www.architecture2030.org/ Boelman EC, Sakulpipatsin P (2004) Critical analysis of exergy efficiency definitions applicable to buildings and building services. In: Plea 2004 – the 21st conference on passive and low energy architecture, Eindhoven, 19–22 Sept 2004 Hartkopf V et al (2002) Building as power plant. The Austin papers. Building Green,. ISBN 1-929884-10-9 Loftness V (2000) Energy, productivity and the critical role of the built environment. In: RAND/DOE E-vision workshop proceedings, Arlington, 11–13 Oct 2000 United Nations Sustainable Development Goals. Sustainable Development Goal 11: make cities and human settlements inclusive, safe, resilient and sustainable. https://sustainabledevelopment.un.org/sdg11

Part I Models of Sustainable Built Environments

Bioclimatic Design Donald Watson FAIA, Trumbull, CT, USA

Article Outline Glossary Definition of the Subject Introduction Principles of Bioclimatic Design Practices of Bioclimatic Design Bioclimatic Design of Atriums Large-Scale Applications Urban and Regional Scale Future Directions: Design for Resilience to Climate Change Bibliography

Glossary Terms and symbols frequently used in building science and climatology Celsius temperature ( C) refers to temperatures measured on a scale devised in 1742 by Anders Celsius, a Swedish astronomer. The Celsius scale is graduated into 100 units between the freezing temperature of water (0  C) and its boiling point at normal atmospheric pressure (100  C) and is, consequently, commonly referred to as the centigrade scale. Dew point temperature (DPT) is the temperature of a surface upon which water vapor contained in the air will condense into liquid water. Stated differently, it is the temperature at which a given quantity of air will become saturated (reach 100% relative humidity) if chilled at constant pressure. It is thus another indicator of the moisture content of the air. Dew point temperature is not easily measured directly; it is conveniently found on a

psychrometric chart if dry-bulb and wet-bulb temperatures are known. Dry-bulb temperature (DBT) is the temperature measured in air by an ordinary (dry bulb) thermometer and is independent of the moisture content of the air. It is also called “sensible temperature.” Fahrenheit temperature ( F) refers to the temperature measured on a scale devised by G. D. Fahrenheit, the inventor of alcohol and mercury thermometers, in the early eighteenth century. On the Fahrenheit scale, the freezing point of water is 32  F, and its boiling point is 212  F at normal atmospheric pressure. Humidity is a general term referring to the water vapor contained in the air. Like the word “temperature,” the type of “humidity” must be defined. Relative humidity (RH) is defined as the percent of moisture contained in the air under specified conditions compared to the amount of moisture contained in the air at total saturation at the same (dry bulb) temperature. Relative humidity can be computed as the ratio of existing vapor pressure to vapor pressure at saturation or the ratio of absolute humidity to absolute humidity at saturation existing at the same temperature and barometric pressure. Wet-bulb temperature (WBT) is an indicator of the total heat content (or enthalpy) of the air, that is, of its combined sensible and latent heats. It is the temperature measured by a thermometer having a wetted sleeve over the bulb from which water can evaporate freely.

Definition of the Subject Bioclimatic design – combining “biology” and “climate” – is an approach to the design of buildings and landscape that is based on local climate. Bioclimatic design techniques include solar heating and sun shading, natural ventilation, and use of building materials for thermal time lag and storage.

© Springer Science+Business Media, LLC, part of Springer Nature 2020 V. Loftness (ed.), Sustainable Built Environments, https://doi.org/10.1007/978-1-0716-0684-1_225 Originally published in R. A. Meyers (ed.), Encyclopedia of Sustainability Science and Technology, © Springer Science+Business Media LLC 2017 https://doi.org/10.1007/978-1-4939-2493-6_225-3

19

20

Resilient design is an extension of bioclimatic design, adding precautionary measures to provide health and safety to prepare buildings, communities, and regions for natural disasters and climate change.

Introduction In adopting bioclimatic approaches, the designer endeavors to create comfort conditions in buildings by understanding the microclimate and resulting design strategies that include natural ventilation, daylighting, and passive heating and cooling. Examples of bioclimatic designs are found in examples of indigenous and vernacular building throughout the globe, evidence of genius loci, the ways of living and working rooted in a particular place and time. Now an established topic of building science research and architectural practice, bioclimatic design can be applied to buildings, landscapes, and urban and regional scales, as part of the twenty-first century sustainability and resilient planning goals. Techniques of sun-tempering, solar shading, and daylighting were amply represented in the early twentieth century portfolios of Frank Lloyd Wright, Tony Garnier, and Augustin Rey, in the 1920s Bauhaus manifestos of Hannes Meyer and Marcel Breuer, in the late 1920s health-oriented designs of Alvar Aalto and Richard Neutra, and in “solar house” designs of the Keck brothers in Chicago area in the late 1930s [1]. Olgyay and Olgyay used the term “bioclimatic design” to define a methodology that matched local climate variables to the achievement of human comfort, applicable to architecture and planning, promoted in a series of professional and popular publications in the late 1940s and 1950s [2, 3]. When air-conditioning systems became widely available in the 1950s and electricity was considered cheap and available, it became possible “to cool a glass house in the desert.” Interest in bioclimatic design waned and became less evident in built work, although pioneering studies continued in academic and architectural research centers in

Bioclimatic Design

Britain and the United States. The topic revived in response to energy shortages of the 1970s. “Passive solar design” became the popular term to incorporate elements of bioclimatic design, at first emphasizing solar heating but soon broadened to passive cooling and daylighting [4]. In the late 1980s, the United Nations Bruntland Commission and the Rio Earth Summit of June 1992 gave international focus to the concept of “sustainability,” including reduced reliance on nonrenewable resources and threats to ecosystems and the biodiversity of environments and cultures. With emergence of such global environmental concerns, the scope of bioclimatic design was enlarged to include landscape, soil, water, and waste nutrient recovery, designed to mimic and restore the health of natural processes and ecosystems, characterized by the term “regenerative design” [5]. At the beginning of the twenty-first century, the world confronts the evident trends of extreme weather and climate change. Bioclimatic design has gained additional relevance as the basis of applying climate science to “passive survivability” – defined by Alex Wilson as “a building’s ability to maintain critical life-support conditions if services such as power, heating fuel, or water are lost” [6]. The challenge to reduce and eliminate where possible the use of fossil fuels for carbon reduction further supports bioclimatic design. Bioclimatic design provides the knowledge and inspiration of nature to design for sustainability and resilience in buildings, landscapes, cities, and regions [7].

Principles of Bioclimatic Design Six key variables have been identified in research studies of human physiological comfort, in which volunteer subjects are asked to report level of comfort and discomfort: air temperature, ambient radiant temperature, humidity, air velocity, dress, and metabolic rate as a function of activity level [8]. Common responses across all these variables have established baseline conditions required for

Bioclimatic Design

human comfort. Studies indicate that the “comfort zone” for humans does not vary, regardless of sex, age, place of origin and residence, skin color, and body form and weight. In other words, the human “comfort” zone is relatively universal independent of age, health, or sex. However, points where research subjects report discomfort and evidence thermal stress do vary, as a function of many variables, including age, health, and acculturation. Comfort studies are the basis of the design criteria of heating and cooling systems for buildings, also applicable to bioclimatic design. The “resources” of bioclimatic design are the natural flows of “ambient” energy in and around a building – the “microclimate” created by the sun, wind, precipitation, vegetation, and temperature and humidity in the air and in the ground (Fig. 1). • Conduction – from hotter object to cooler object by direct contact. • Convection – by flow of air between warmer objects and cooler objects. • Radiation – from hotter object to cooler object within the direct view of each other regardless

Bioclimatic Design, Fig. 1 Paths of energy exchange at the building microclimate scale. Bioclimatic design is based upon understanding energy flows within and around buildings (Ref. [4])

21

of the temperature of air between, including radiation from the sun to the earth. • Evaporation – the change of phase from liquid to gaseous state: the sensible heat (dry-bulb temperature) in the air is lowered by the latent heat absorbed from air when moisture is evaporated. • Thermal storage – from heat charge and discharge both diurnally and seasonally, as a function of its specific heat, mass, and conductivity. Although not usually included alongside the four classic means of heat transport, thermal storage is helpful in understanding the heat transfer physics of building climatology. The strategies can be set forth as: • Minimize conductive heat flow This strategy is achieved by using insulation and thermal breaks. It is effective when the outdoor temperature is significantly different, either lower or higher, than the interior comfort range. In summer, this strategy should be considered whenever ambient temperatures are within or above the comfort range and where natural cooling strategies cannot be relied upon to achieve comfort (i.e., whenever mechanical air conditioning is necessary). • Delay periodic heat flow While the insulation value of building materials is well understood, it is not widely appreciated that building envelope materials also can delay heat flows that can be used to improve comfort and to lower energy costs. Time lag through masonry walls, for example, can delay the day’s thermal impact until evening and is a particularly valuable technique in hot arid climates with wide day-night temperature variations. Techniques of earth sheltering and earth berming also exploit the long-lag effect of subsurface construction. • Minimize infiltration “Infiltration” refers to uncontrolled air leakage around doors and windows and through joints, cracks, and faulty seals in the building envelope. Infiltration (and resulting “exfiltration” of heated or cooled air) is considered the largest and potentially the most

22













Bioclimatic Design

intractable source of energy loss in a building, once other practical measures have been taken. Provide thermal storage Thermal mass inside the insulated envelope is critical to dampening of the swings in air temperature and in storing heat in winter and as a heat sink in summer. Promote solar gain The sun can provide a substantial portion of winter heating energy through elements such as equatorial-facing windows and greenhouses that include other passive solar techniques which utilize spaces to collect, store, and transfer solar heat. Minimize solar gain The best means for ensuring comfort from the heat of summer is to minimize the effects of the direct sun by shading windows from the sun, or otherwise minimizing the building surfaces exposed to summer sun, by use of radiant barrier, and by insulation. Minimize external airflow Winter winds increase the rate of heat loss from a building by “washing away” heat and thus accelerating the cooling of the exterior envelope surfaces by conduction and also by increasing infiltration (exfiltration) losses. Siting and shaping a building to minimize wind exposure or providing windbreaks can reduce wind impacts and heat loss. Promote ventilation Cooling by airflow through an interior may be propelled by two natural processes, crossventilation (wind driven) and stack-effect ventilation (driven by the buoyancy of heated air even in the absence of external wind pressure). A fan (using photovoltaic for fan power) can be an efficient way to augment natural ventilation cooling in the absence of sufficient wind or stack-pressure differential. Promote radiant cooling A building can lose heat if the mean radiant temperature of the materials at its outer surfaces is greater than that of its surroundings, principally the night sky. The mean radiant temperature of the building surface is determined by the intensity of solar irradiation, by the material surface (film coefficient), and by

the emissivity of its exterior surface (its ability to “emit” or reradiate heat). This contributes only marginally if the building envelope is well insulated. • Promote evaporative cooling Sensible cooling of a building interior can be achieved by evaporating moisture into the incoming airstream (or if an existing roof has little insulation, by evaporative cooling the exterior envelope such as by a roof spray.) These are simple and traditional techniques and most useful in hot dry climates if water is available for controlled usage. Mechanically assisted evaporative cooling is achieved with an economizer-cycle evaporative cooling system, instead of, or in conjunction with, refrigerant air conditioning. The “comfort range” as defined in research studies is within a small range of temperature and humidity conditions, roughly between 68–80  F (20–26.7  C) and 20–80% relative humidity (RH), referred to on the psychrometric chart as the “comfort zone.” Other variables include environmental indices – radiant temperature and rate of airflow – as well as clothing and activity (metabolic rate). While such criteria describe relatively universal requirements in which all humans are “comfortable,” there are significant differences in and varying tolerance for discomfort, that is, the limits in which stress is felt, which vary depending upon age, sex, health, cultural conditioning, and expectations. Givoni [9] and Milne and Givoni [10] offer a design method using the “Building Bioclimatic Chart,” modified by Arens [11]. (Fig. 2) Adopting the psychometric chart format, the Building Bioclimatic Chart displays the parameters for bioclimatic design strategies that can achieve human comfort in a building interior. If local outdoor temperatures and humidity fall within specified zones, the designer is alerted to opportunities to use specific bioclimatic design strategies to create effective interior comfort. Typical meteorological year (TMY) summaries contain climatic data for all 8,760 h in a “typical” year, available for most locations in the United States and increasingly available for major

Bioclimatic Design

23

Bioclimatic Design, Fig. 2 Building Bioclimatic Chart. The chart indicates parameters of climatic conditions favorable for bioclimatic design (Refs. [9, 10, 11])

regions and cities worldwide. Each file contains one complete year of hourly data, including direct (beam) solar radiation, total horizontal solar radiation, dry-bulb temperature, dew point humidity, wind speed, and cloud cover [12]. Climate Consultant is a computer-based program that can be downloaded at no cost from the web [13]. Part of a career-long project of UCLA Professor Emeritus Murray Milne to develop public domain energy design tools, the software plots temperatures, wind velocity, sky cover, percent sunshine, beam, and horizontal irradiation. It includes 3-D plots of temperature, wind speed, and related climatic data cross-referenced to bioclimatic design practices presented in Watson and Labs [4]. Climate Consultant graphs include the Building Bioclimatic Chart for Atlanta (Fig. 3), with summaries of percent annual hours of heating and cooling requirements, along with effective bioclimatic strategies, listed below in rank order of

percent annual hours of potential need and effectiveness. The designer can thus assess the relative effectiveness and priorities of options, also subject to local energy costs, reliability, and building uses: 26.8% Heating, add humidification if needed (2346 h) 25.4% Internal heat gain (2223 h) 17.2% Dehumidification (1504 h) 14.0% Sun shading of windows (1228 h) 11.2% Cooling, add dehumidification if needed (979 h) 11.1% Comfort (968 h) 09.9% Passive solar direct gain low mass (866 h) 08.5% Passive solar direct gain high mass (747 h) 03.4% High thermal mass (299 h) 02.9% High thermal mass night flushed (252 h) 02.3% Two-stage evaporative cooling (198 h) 02.2% Direct evaporative cooling (189 h) 02.0% Natural ventilation cooling (174 h)

24

Bioclimatic Design

Bioclimatic Design, Fig. 3 Climate Consultant display of the Building Bioclimatic Chart for Atlanta GA. Representative bioclimatic chart generated by Climate

Consultant. The box “Design Strategies” tabulates the percent hours per year that bioclimatic design strategies are effective, compiling TMY3 data set (Ref. [13])

01.2% Fan-forced ventilation cooling (107 h) 01.1% Wind protection of outdoor spaces (94 h) 00.0% Humidification only (0 h)

Thermal Envelope (Winter) Isolating the interior space from the hot summer and cold winter climate, such as:

Practices of Bioclimatic Design Bioclimatic techniques can be set forth as a set of design opportunities, adapted from Watson and Labs [4]. Windbreaks (Winter) Two design techniques serve the function of minimizing winter wind exposure: • Use neighboring landforms, structures, or vegetation for winter wind protection. • Shape and orient the building shell to minimize winter wind turbulence (Fig. 4).

• Use attic space as buffer zone between interior and outside climate. • Use basement or crawl space as buffer zone between interior and grounds. • Use vestibule or exterior “wind shield” at entryways. • Locate low-use spaces, storage, utility, and garage areas to provide climatic buffers. • Subdivide interior to create separate heating and cooling zones. • Select insulating materials for resistance to heat flow through building envelope. • Apply vapor barriers to the warm side of building envelope assemblies to control moisture migration.

Bioclimatic Design

25

Bioclimatic Design, Fig. 4 Sea Ranch, California. Landscape planting, roof slopes, and fencing designed for wind protection. Esherick, Homsey, Dodge, and Davis, architects and planners, with Lawrence Halprin, landscape architect

Bioclimatic Design, Fig. 5 Analysis of building aspect ratio. Simplified building shapes are compared for ratio of exterior surface to enclosed volume (Ref. [4])

• Develop construction details to minimize air infiltration and exfiltration. • Provide insulating controls at glazing. • Use heat reflective or radiant barriers on or below surfaces oriented to summer sun. • Minimize the outside wall and roof area ratio of exterior surface to enclosed volume (Fig. 5).

Solar Windows and Walls (Winter) Using the winter sun for heating a building through solar-oriented windows and walls is provided by a number of techniques: • Maximize reflectivity of ground and building surfaces outside windows facing the winter sun.

26

Bioclimatic Design

Bioclimatic Design, Fig. 6 Solar windows and walls. Left: Keck and Keck, architects, developed solar design principles in Chicago area in the 1930s. The designs of Keck and Keck – in this example a prefab for green readybuilt homes – included south-facing glass, exposed masonry floors with hypostyle (warm air radiant) heating,

interior masonry walls, interior curtains, and exterior shading (Photo: William Keck, architect). Right: Rural school, Kasserine, Tunisia. Solar “Trombe Wall” (glass-covered masonry), clerestory daylighting, Save The Children Federation. 1986 (Photo: Donald Watson, FAIA, architect)

• Shape and orient the building shell to maximize exposure to winter sun. • Use high-capacitance thermal mass materials in the interior to store solar heat gain. • Use solar wall and roof collectors on equatorial-oriented surfaces. • Optimize the area of equatorial-facing glazing. • Use clerestory skylights for winter solar gain and natural illumination. • Provide solar-oriented interior zone for solar heat gain, with solar control for shading in overheated periods (Fig. 6).

• Use slab-on-grade construction for ground temperature heat exchange and thermal storage. • Use earth-covered or sod roofs. • Recess structure below grade or raise existing grade for earth sheltering (Fig. 8).

Indoor/Outdoor Rooms (Winter and Summer) Courtyards, covered patios, seasonal screened and glassed-in porches, greenhouses, atriums, and sun spaces can be located in the building plan for summer cooling and winter heating benefits, as in these three techniques: • Provide outdoor semi-protected areas for yearround climate moderation (Fig. 7). Earth Sheltering (Winter and Summer) Techniques such as banking earth against the walls of a building or green roofs provide thermal storage and damping temperature fluctuations (daily and seasonally), reducing envelope heat loss or gain (winter and summer). These techniques are often referred to as earth contact or earth sheltering:

Thermally Massive Construction (Summer and Winter) Particularly effective in hot arid zones or in more temperate zones with cold clear winters. Thermally massive construction provides a “thermal fly wheel.” Absorbing heat during the day from solar radiation and convection from indoor air, thermal mass can create comfort if it is cooled at night, if necessary through nighttime ventilative cooling (if air temperatures fall within the comfort zone): • Use high mass construction with outside insulation and nighttime ventilation. • For selected climates (hot dry), use highcapacitance materials to dampen heat flow through the building envelope (Fig. 9). Sun Shading (Summer) Mid-day solar altitude angles are higher in summer than in winter. Thus, an overhang can shade windows from the sun during the overheated summer period and permit sun to reach the window surfaces and interior spaces in winter:

Bioclimatic Design

27

Bioclimatic Design, Fig. 7 Protected courtyard. Buli Khelam Lhakhang Monastery, Bhutan. In the Himalayan tradition of building, an enclosed courtyard with sunexposed adobe walls and windows creates a wind-

protected microclimate, permitting a temperate planting regime to flourish within, in contrast to high mountain climatic conditions of its locale (Photo: Donald Watson)

• Minimize reflectivity of ground and building surfaces outside windows facing the summer sun. • Use neighboring landforms, structures, or vegetation for shading summer sun. • Shape and orient the building shell to minimize exposure to summer afternoon sun. • Provide seasonally operable shading, including deciduous trees.

• Orient door and window openings to facilitate natural ventilation from prevailing summer breezes. • Use wing walls, overhangs, and louvers to direct summer wind flow into interior. • Use louvered wall openings for maximum ventilation control. • Use roof monitors for “stack effect” ventilation (Fig. 10).

Natural Ventilation (Summer and Seasonal) Natural ventilation is a simple concept by which to cool a building:

Plants and Water (Summer) Many techniques provide cooling by plants and water near building surfaces for shading and evaporative cooling:

• Shape and orient the building shell to maximize exposure to summer breezes. • Use “open plan” interior to promote airflow. • Provide vertical airshafts to promote “thermal chimney” or stack-effect airflow. • Use double roof construction for ventilation within the building shell.

• Use planting next to building skin (provided it does not interfere with ventilation). • Use roof spray or roof ponds for evaporative cooling. • Use ground cover and planting for site cooling. • Maximize on-site evaporative cooling (Fig. 11).

28

Bioclimatic Design

Bioclimatic Design, Fig. 8 Earth-sheltered home. New Canaan, CT. Donald Watson, FAIA, architect. (1) Winter solstice, (2) summer solstice, (3) light shelf/daylight reflector, (4) green roof, (5) skylight, (6) earth sheltering

Bioclimatic Design, Fig. 9 Thermal mass appropriate for hot dry climate. Indigenous adobe block construction, with roof and window overhangs to shade and protect the walls. Tahono O’Odham Nation, Papago Indian Reservation, Arizona (Photo: Donald Watson)

Bioclimatic Design of Atriums Atriums offer many energy design opportunities, depending upon climate variables, to provide natural heating, cooling, lighting, and plants. Suggested by its Latin meaning as “heart,” or an open courtyard of a Roman house, the term atrium as used today describes a protected courtyard or glazed large-volume space placed within a building. Modern atrium design incorporates many architectural elements – wall enclosures, sunoriented openings, shading and ventilation devices, and subtle means of modifying temperature and humidity – suggested by examples that

derive from nineteenth century greenhouses and glass-covered arcades of Great Britain and France. In Northern Europe, especially Holland and England, from the seventeenth century onward, south-facing orientation of indoor gardens, propagating sheds, orangeries, and conservatories revealed an understanding of bioclimatic design. Gardeners and greenhouse designers combined thermal mass, double glass, steep glass orientation, underground heating, shading, and insulating devices in greenhouses. The greenhouse designs of J.C. Loudon, beginning circa 1820, had all of these elements evident in sketches and

Bioclimatic Design

29

Bioclimatic Design, Fig. 10 Shading and ventilation strategies. Built in an era well before air-conditioning, plantation manor houses such as the 1827 San Francisco Plantation House, New Orleans, combined a range of

strategies for natural cooling in hot humid climates, including open understory and porches, cross-ventilation, and roofs designed to induce ventilation by thermal updraft (Photo: Robert Perron)

built examples through mid-century. Joseph Paxton’s The Great Exhibition or the Crystal Palace Exhibition of 1851 demonstrated the possibility for large glazed-covered areas, inaugurating a proliferation of urban atrium designs across Europe and the world [14]. Atriums offer many energy design opportunities: first, comfort is achieved by gradual transition from outside climate to building interior; second, designed properly, protected spaces and buffer zones create natural and free-flowing energy by reducing or by eliminating the need to otherwise heat, cool, or light building interiors.

vertical or sloped not lower than a tilt angle equal to the local latitude. Heating Rule 2 To store and distribute heat, place interior masonry directly in the path of the winter sun. This is most useful if the heated wall or floor surface will in turn directly radiate to building occupants. Heating Rule 3 To prevent excessive nighttime heat loss, consider an insulating system for the glazing, such as insulating curtains or highperformance multilayered window systems. Heating Rule 4 Heat recovery can be accomplished if the warm air is redistributed either to the lower area of the atrium (a ceiling fan) or redirected (and cleaned) to the mechanical system or through a heat exchanger if the air must be exhausted for health and air-quality reasons. Because a large air volume must be heated, an atrium is not an efficient solar collector. A high space helps to make an overheated

Solar Heating Guidelines If heating efficiency alone is the primary energy design goal of the atrium, the following design principles should be paramount: Heating Rule 1 To maximize winter solar heat gain, orient the atrium aperture (openings and glazing) to the equator. If possible, the glazing should be

30

Bioclimatic Design

Natural Cooling Guidelines Several guidelines related to the use of an atrium design as an intermediary or buffer zone apply to both heating and cooling. If an unconditioned atrium is located in a building interior, heat gain results from the warmer surrounding spaces into the atrium. In buildings with large internal gains due to occupants, lighting, and machines, the atrium may require cooling throughout the year. To design exclusively for cooling, the following principles would predominate:

Bioclimatic Design, Fig. 11 Evaporative cooling strategies: Public courtyard. Seville, Spain. The streets and passages of the city combine courtyards, gardens, and a landscape rich with planting and water fountains (Photo: Helen Kessler)

space acceptable, as the warmest air rises to the top. By facing a large skylight and/or window opening toward the equator, direct winter solar heating becomes feasible. In cool climates, an atrium used as a solar heat collector would require as much winter sunlight as possible. In overbright conditions, dark finishes on surfaces where the sun strikes will help reduce glare and also to store heat. On surfaces not in direct sun, light finishes reflect light, especially welcomed under cloudy conditions. In most locations and uses, glass should be completely shaded from the summer sun. Movable insulation might be considered to reduce nighttime heat loss.

Cooling Rule 1: To minimize solar gain, provide shade for the summer sun. While fixed shading devices suffice for much of the summer period, movable shading is the only means by which to match the seasonal shading requirements at all times. In buildings in warm climates, sun shading may be needed throughout the year. Cooling Rule 2: To use the atrium as an exhaust air plenum in the mechanical system of the building. The great advantage is one of economy, but heat recovery options (discussed above) and ventilation become most effective when the natural airflow in the atrium is in the same direction and integrated with the mechanical system. Cooling Rule 3: To facilitate natural ventilation, create a vertical “chimney” effect by placing ventilating outlets high (preferably in the freeflow airstream well above the roof) and by providing cool “replacement air” inlets at the atrium bottom, with attention that the airstream is clean, that is, free of car exhaust or other pollutants. The inlet air steam can be cooled naturally, best with cool air from a shaded area. In hot, dry climates, passing the inlet air over water such as an aerated fountain or landscape can facilitate evaporative cooling. Allowing the atrium to cool by ventilation at night is effective in climates where summer nighttime temperatures are lower than daytime (greater than 15  F difference). Additional cooling capacity (to absorb and hold heat) is provided by materials such as masonry. However, as a general rule, if the average daily temperature is above 78  F (25.5  C),

Bioclimatic Design

thermally massive materials are disadvantageous in non-air-conditioned spaces because they do not cool as rapidly as a thermally light structure. When stack ventilation is possible through a roof aperture, the space will ventilate naturally even in the absence of outside breezes, by the driving force of heated air. If air conditioning of the atrium is needed but can be restricted to the lower area of the space, it can be done reasonably; cool air, being heavier, will pool at the bottom. Design choices must balance between the requirements for sun shading and those for daylighting. The ideal location for a shading screen is on the outside of the glazing, where it can be wind-cooled. When the outside air ranges about 80F (26.7  C), glass areas – even if shaded – admit undesired heat gain by conduction. In truly warm climates, a minimum of glazed aperture should be used to prevent undesired heat gain: a small amount of glazing should be placed where it is most effective for daylighting. Heat-absorbent or heat-reflective glass, the common solution to reduce solar heat gain, reduces the illumination level and also reduces desirable winter heat gain. In temperate-to-cool climates, heat gain through a skylight can be tolerated if the space is high, so that heat builds up well above the occupancy zone and there is good ventilation. In hot climates, an atrium will perform better as an unconditioned space if it is a shaded but otherwise open courtyard. Daylighting Guidelines In all climates, an atrium can be used for daylighting. Electric lighting cost savings can be achieved but only if the daylighting system works, that is, if it replaces the use of artificial lighting. (Many daylit buildings end up with the electric lights in full use regardless of lighting levels needed.) Atriums serve a particularly useful function for an entire building by balancing light levels – thus reducing brightness ratios – across the interior floors of a building. If, for example, an open office floor has a window wall on only one side, typically more electric lighting is required than would be required without natural lighting to reduce the brightness ratio. A light court can provide such balanced “two-source” lighting.

31

The following principles apply to atrium design for daylighting: Lighting Rule 1 To maximize daylight, an atrium cross section should be stepped open to the entire sky dome in predominantly cloudy areas. In predominantly sunny sites, atrium geometry can by based upon heating and/or cooling solar orientation principles. Lighting Rule 2 To maximize light, window or skylight apertures should be designed for the predominant sky condition. If the predominant sky condition is cloudy and maximum daylight is required (as in a northern climate winter garden), consider clear glazing oriented to the entire sky dome, with movable sun controls for sunny conditions. If the predominant sky condition is sunny, orient the glazing according to heating and/or cooling design requirements. Lighting Rule 3 Provide sun-and-glare control by geometry of aperture, surface treatment, color, and adjustable shades or curtains. Designing for daylighting involves compromise to meet widely varying sky conditions. What works in bright sun conditions will not be adequate for cloudy conditions. An opaque overhang or louver, for example, may create particularly somber shadowing on a cloudy day. Light is diffused by a cloudy sky, falling nearly equally from all directions; the sides of the atrium thus cast gray shadows on all sides. For predominantly cloudy conditions, a clear skylight is the right choice. Bright haze will nonetheless cause intolerable glare at least to a view upward. Under sunny conditions, the same skylight is the least satisfactory choice because of overlighting and overheating. Unless the local climate is truly cloudy and the atrium requires high levels of illumination, partial skylighting can achieve a balance of natural lighting, heating, and cooling. Partial skylighting is a skylighting that takes only a portion of the roof surface. This approach offers advantages of controlling glare and sunlight by providing reflecting

32

and shading surfaces, such as by the coffers of the skylights. With less light intensity and contrast, a surface illuminated by reflected light is more acceptable to the human eye than a direct view of a bright window area. Movable shades for glare and sun control provide a further means to balance for the variety of conditions. The design principles for heating, cooling, and daylighting can be selected according to building type and local climate. In northern climates, the solar heating potential predominates, while the natural cooling potential predominates in the Southern United States. In commercial and institutional structures, natural cooling and daylighting are both important. In this case, the local climate would determine the relative importance of openness achieved with large and clear skylighting (most appropriate for cloudy temperate-to-cool regions) or of closed and shaded skylighting (most appropriate for sunny warm regions). The design principles can be summarized as guideline principles (Fig. 12). Garden Atriums Plants have an important role in buffer zones. If the requirements of plants are understood, healthy greenery can be incorporated into atrium design and contribute to human comfort, amenity, and energy conservation. Plants, however, when uncomfortable, cannot move. Major planting losses have been reported in gardened atriums because the bioclimatic requirements were not achieved. A greenhouse for year-round crop or plant production is intended to create springsummer or the growing period climate throughout the year. A winter garden replicates springsummer conditions for plant growth in wintertime by maximizing winter daylight exposure and by solar heating. Plants need ample light, but not excessive heat. Although varying according to plant species, as a general rule, planting areas require full overhead skylighting (essentially to simulate their indigenous growing condition). Most plants are overheated if their roots range above 65  F (18.3  C). Plant growth slows when the root temperature drops below 45  F (7.2  C). As a result, a greenhouse has the general problem of overheating (as well as overlighting) during

Bioclimatic Design

any sunny day and of underlighting (in intensity and duration) during any cloudy winter day. If the function of the atrium includes plant propagation or horticultural exhibit (replicating the indigenous climate in which the display plants flower), then clear-glass skylighting is needed for the cloudy days, and adjustable shading and overheating controls are needed for sunny days. If the plant beds are heated directly, for example, by water piping, then root temperatures can be maintained in the optimum range without heating the air. As a result, the air temperature in the atrium can be cool for people, in the 50  F (10  C) range, with the resulting advantage of providing a defense against superheating the space. People can be comfortable in lower air temperatures if exposed to the radiant warmth of the sun and/or if the radiant temperature of surrounding surfaces is correspondingly higher, that is, ranging above 80  F (26.7  C). Lower atrium temperatures have a further advantage to plants and energy-efficient space operation: evaporation from plants is slowed, saving water and energy (1000 Btu is removed from the sensible heat of the space with each pound of water that evaporates). Air circulation reduces excessive moisture buildup at the plant leaf and circulates CO2, needed during the daytime growth cycle (Fig. 13).

Large-Scale Applications Figure 14 depicts site and building opportunities for energy collection, storage, and distribution that may be integrated into larger buildings as combined passive and active measures of bioclimatic design. Applications may include on-site ecosystem services, green roofs, water collection, waste recycling, and biological diversity of sun and shade, warm and cool zones, and energy storage, selected depending upon opportunities within each site and region. William Lam [16] provides a detailed guidance for sunlighting large buildings, including documentation of case studies. A number of largescale building designs demonstrate exemplary applications of microclimatic design (Figs. 15, 16, and 17).

Bioclimatic Design

33

Bioclimatic Design, Fig. 12 Bioclimatic principles for atrium design. Guidelines for design of atria and light courts in various climates (Ref. [15])

34

Bioclimatic Design, Fig. 13 New Canaan Nature Center Greenhouse, New Canaan, CT 1982. Buchanan and Watson, architects. (1) South-facing greenhouse, (2) solar collectors, (3) thermal storage, (4) ceiling fans, (5) roof

Bioclimatic Design

monitor, (6) operable sun shade/insulating curtain, (7) earth-contact floor, (8) root-bed heating (9) grow lights, (10) earth berm, (11) rainwater collection (Illustration: Marja Watson)

Bioclimatic Design, Fig. 14 Large building opportunities for microclimatic design integration. Bioclimatic design extends to larger scale for integrated heating, cooling and lighting systems

Urban and Regional Scale Many studies address microclimatic impacts at the urban scale [3]. Perennial topics have included solar access, evident in early twentieth century studies related to solar access and daylighting, as well as urban scale airflow, Solar Access Ralph Knowles [17] in studies undertaken over many decades with students at the University of

Southern California developed the notion of assuring solar access to buildings, for sun tempering, daylighting, and solar collection. His studies have demonstrated that solar access can be guaranteed in most urban areas while keeping within conventional medium- to medium-highdensity floor to area ratios (FARs) (Fig. 17). A study by FXFOWLE Architects [18] illustrates the feasibility of passive solar and improved insulation measures equal to Passivhaus standards. Strategies include shading, passive solar gain, shading,

Bioclimatic Design

Bioclimatic Design, Fig. 15 Centre for Interactive Research on Sustainability (CIRS), University of Endowment Lands, British Columbia, Vancouver. Designed as a “living laboratory” with multiple innovations to reduce energy and to capture and use rainwater. Busby Perkins+Will, architects. (1) Daylighting/sun tempering, (2) photovoltaic collectors, (3) evacuated tube collectors, (4) green

35

roof/rainwater harvesting, (5) displacement ventilation, (6) ground source heat pump, (7) heat recovery, (8) radiant heating, (9) deciduous living wall, (10) solar aquatic biofiltration, (11) stormwater to raingardens, (12) solar DHW, (13) rainwater cistern, (14) water purification, (15) gray and blackwater recovery (Photo: Martin Tessler)

Bioclimatic Design, Fig. 16 Solaire – 27-story residential apartment building in New York City. Passive solar, green roof, energy- and water-conserving features. Cesar Pelli & Associates Architects and SLCE Architects

36

Bioclimatic Design

Bioclimatic Design, Fig. 17 Left: Solar envelope for a medium-density neighborhood of Los Angeles. Right: Mediumdensity neighborhood within the solar envelope (Photos: courtesy of Ralph L. Knowles)

Bioclimatic Design, Fig. 18 Case study of Passivhaus standards applicable with 2016 New York City zoning and housing market requirements. Buildings in color indicate

“full build out” of surrounding properties, with winter morning shading (Ref. [18])

attention to thermal breaks in insulation and envelope, sun-oriented interior layouts, and energyrecovery within mechanical ventilation (Fig. 18).

Table 1 shows averages of air and surface temperatures measured at a height of 1 m (3.3 ft.) around noontime on the UCLA campus during a sequence of several clear days in summer. The lowest temperatures were in a space between a line of high shrubs and a wall of a building. Givoni’s research points to opportunities for continued research at the urban scale, supporting an approach to urban planning based on bioclimatic analysis and design. (Fig. 19)

Urban Heat Islands and Cool Zones Baruch Givoni [18] compiles a broad survey of urban bioclimatic data and design applications, with emphasis on measured data, along with discussion of challenges of data measurement at the urban scale.

Bioclimatic Design

37

Bioclimatic Design, Table 1 Representative air and surface temperature averages measured during a sequence of several clear days in summer (Ref. [18]) Location Parking lot Open plaza Shaded walk Grass lawn Behind shrubs

Air temperature ( F) 79 78 76 75 74

Surface temperature ( F) 122 107 80 88 73

Air temperature ( C) 26.1 25.6 24.4 23.9 23.3

Surface temperature ( C) 50.0 41.7 26.7 31.1 22.8

Bioclimatic Design, Fig. 19 Pocket Park, New York City. Paley Park creates a small area of respite, with a cooling microclimate created by evaporative cooling, shading, and wind protection, while water fountain sound helps neutralize urban clamor (Photo: Donald Watson)

Urban Air Quality Studies by Anne Whiston Spirn [19] have utilized research on urban wind effects to indicate design strategies to reduce pollution in city streets and public ways, principally by opening building forms and landscapes to less constrained airflow (Figs. 20 and 21).

Future Directions: Design for Resilience to Climate Change The concept of resiliency applies lessons from natural systems to design for safety and protection in extreme conditions using strategies found in

natural systems, such as buffering, zone separation, redundancy, rapid feedback, and decentralization. Extreme conditions include impacts of natural disasters, such as hurricanes, tsunamis, and earthquakes. It also includes mitigation and adaptation measures for longer-term risks of global warming and sea level rise through actions that reduce carbon emissions. As cities grow in size and density, risks to life safety and health increase. The natural landscape that has evolved in response to climate and water regimes over millennia had adapted to long-evolving patterns of rainfall, aridity, heat, and cold. Historical flood conditions were accommodated within the

38

Bioclimatic Design

Bioclimatic Design, Fig. 20 Air quality. The urban designer has opportunity to utilize strategies to improve air quality at the urban microclimatic scale (Ref. [19]). (a) Street canyons lined with building of similar height, oriented perpendicular to the wind direction tend to have poor air circulation compared to (b). (b) Street canyons

lined with buildings of different heights and interspersed with open areas have better air circulation. (c) To promote air circulation in street canyons, step buildings back from the street, increase openings and vary building heights. (d) To promote air circulation in street side arcades, design them with high canopies and airflow outlets

watershed ecology and its coevolving plants and animals. When those patterns are disrupted and the natural landscape is altered, flooding risks and disasters increase. Watson and Adams [7] and Watson [21] extend bioclimatic design to include resilience, to adopt precautionary principles in design of buildings, communities, and cities. Resiliency is evident in natural systems strategies to adjust to shock, variable, and extreme conditions (Table 2). Bioclimatic lifeline systems – green space, water, food, waste, mobility, and safe shelter – replicate the biological systems of water, vegetation, food, and biodiversity that protect the life, health, and safety of cities. Ecosystems regulate the supply and quality of water, air, and soil. Urban parks and vegetation reduce the urban heat island effect. Urban green spaces help to regulate climate, reflect and absorb

solar radiation, filter dust, store carbon, serve as windbreaks, improve air quality by oxygen emission and moistening, and enhance cooling by evaporation, shading, and air exchange (Fig. 22). Bioclimatic techniques that contribute to lifeline systems at the urban scale include: Greenspace: walkways, pocket parks, playgrounds, wildlife, trees, plants, and soil protection Water: stream daylighting, cleansing, water fountain cooling zones, and urban wildlife ponds Food: local community gardens, farmer markets, other community market venues Energy: protected utility and communication lines, district energy conduits, solar/wind structures Waste: combined urban services, efficient waste collection, recycling, and removal

Bioclimatic Design

39

Bioclimatic Design, Fig. 21 Comprehensive plan to improve air quality. Stuttgart, Federal Republic of Germany. Public gardens and open space atop the cities’ hills and hillside canyons are preserved as vegetated public

stairways and watercourses. Hillside canyons funnel cool nighttime airflow to center city streets and downtown parks (Photo: Dr. Michael Trieb, Urban Planning Institute, University of Stuttgart)

Bioclimatic Design, Table 2 Lessons of nature for resilient design

Bioclimatic Design, Table 2 (continued)

Principle from nature Absorption

Buffering Core Protection Diffusion Water storage capacity Redundant circuits

Application to resilient design Watershed planning and design (reservoirs, retention ponds, green roofs) Breaks, riparian buffers, rain gardens Zoning, decentralization, self-reliant subsystems Meanders, wetland and coastal zone landscape, open foundations Aquifers, wetlands, reservoirs, cisterns Green infrastructure, wildlife corridors, and multiple service routes (continued)

Principle from nature Waste/nutrient recovery Rapid response

Application to resilient design Sustainable stormwater design and waste systems Early warning, emergencyresponsive systems

Mobility: urban transit options, bikeways, pedestrian-scaled vehicles, flexible use, emergency service lanes Refuge: community shelters and safe zones, emergency communication, and evacuation and materials staging

40

Bioclimatic Design

Bioclimatic Design, Fig. 22 Lifeline systems, integrating bioclimatic principles as urban and regional scales (Ref. [21])

Bibliography 1. Butti K, Perlin J (1980) A golden thread: 2500 years of solar architecture and technology. Van Nostrand Reinhold, New York 2. Fitch JM, Siple P (1952) AIA bulletin 1949–1952. University Microfiche, Ann Arbor 3. Olgyay V, Olgyay A (1963) Design with climate: bioclimatic approach to architectural regionalism. Princeton University Press, Princeton 4. Watson D, Labs K (1993) Climatic building design, 2nd edn. McGraw-Hill, New York 5. Lyle JT (1996) Regenerative design for sustainable development. Wiley, New York 6. Wilson A (2005) Passive survivability: a new design criterion for buildings. Environmental Building News 14:12. www.buildinggreen.com/feature/passive-surviv ability-new-design-criterion-buildings. Accessed 1 June 2017 7. Watson D, Adams M (2011) Design for flooding: architecture, landscape, and urban design for resilience to climate change. Wiley, New York 8. Fanger PO (1970) Thermal comfort. Danish Technical Press, Copenhagen 9. Givoni B (1976) Man, climate and architecture, 2nd edn. Applied Science Publishers, London 10. Milne M, Givoni B (1979) Architectural design based on climate. In: Watson D (ed) Energy conservation

11.

12.

13.

14. 15.

16. 17.

18.

through building design. McGraw Hill, New York, pp 96–113 Arens E, Gonzales R, Berglund L (1986) Thermal comfort under an extended range of environmental conditions. ASHRAE transactions, vol 92-1. ASHRAE Publications, Atlanta, pp 18–25. http:// escholarship.org/uc/item/1jw5z8f2. Accessed 1 June 2017 National Energy Renewable Laboratory NREL (1996) Typical meteorological year climate data files. National Renewable Energy Laboratory, Golden Milne M (1997) Energy design tools. University of California, Los Angeles. www.energy-design-tools. aud.ucla.edu. Accessed 1 June 2017 Hix J (1974) The glass house. MIT Press, Cambridge Watson D (1982) Energy within the space within. Progressive architecture magazine July 1982, pp 97– 102. www.ncmodernist.org/PA/PA-1982-07.PDF. Accessed 1 June 2017 Lam WMC (1986) Sunlighting as formgiver for architecture. Van Nostrand Reinhold, New York Knowles R (2003) The solar envelope. In: Watson D (ed) Time-saver standards for urban design. McGraw-Hill, New York, pp 4.6-1–4.6-18 FXFowle Architects (2017) Feasibility study to implement the Passivehaus standard on tall residential buildings. New York State Research and Development Authority, Albany

Bioclimatic Design 19. Givoni B (2003) Urban design and climate. In: Watson D (ed) Time-saver standards for urban design. McGraw-Hill, New York, pp 4.7-1–4.7-14 20. Spirn AW (2003) Better air quality at street level: strategies of urban design. In: Watson D (ed) Timesaver standards for urban design. McGraw-Hill, New York, pp 7.7-1–7.7-8 21. Watson D (2017) Urban lifelines to archive climate resiliency. In: Bay JWP, Lehmann S (eds) Earthscan/ Routledge, London (forthcoming)

Additional References Brown GZ, DeKay M (2001) Sun, wind & light: architectural design strategies. Wiley, New York Rittelmann BHK, Kantrowitz M (1987) Commercial building design: integrating climate, comfort, and cost. Van Nostrand Reinhold, New York Dodman D, Diep L, Colenbrander S (2017) Resilience and resource efficiency in cities. United Nations Environment Programme, Geneva

41 Fitch JM, Branch DP (1960) Primitive architecture and climate. Sci Am 219(3):190–202 Givoni B (1998) Climate considerations in building and urban design. Van Nostrand Reinhold, New York Hastings SR (ed) (1994) Passive solar commercial and institutional buildings: a sourcebook of examples and design insights. International Energy Agency. Wiley, New York Knowles RL (2006) Ritual houses: drawing on nature’s rhythms for architecture and urban design. Island Press, Washington, DC Koenigsberger OH, Ingersoll TG, Mayhew A, Szokolay SV (1974) Manual of tropical housing and building. Longman, New York Kwok AG, Grondzik WT (2007) The green studio handbook: environmental strategies for schematic design. Elsevier, New York Landsburg HE (1972) Assessment of human bioclimate. Technical Note 123. World Meteorological Organization. UNIPUB, Geneva. http://ac.ciifen.org/omm-biblioteca/ CCI_TECH/WMO-331.pdf. Accessed 1 June 2017

Biophilic Design Catherine O. Ryan and William D. Browning Terrapin Bright Green, New York, NY, USA

Article Outline Glossary Definition of the Subject Introduction Defining Nature Framework Methodology Design Patterns and Implementation Future Directions Bibliography

credible research to achieve the best possible outcomes. Framework Organization or classification of biophilic design patterns in a structured manner with assigned characteristics or guidance. Health A state of complete physical, mental, and social well-being and not merely the absence of disease or infirmity. Nature The living organisms and the nonliving components of an ecosystem. Pattern A description of a problem that occurs repeatedly in the environment, coupled with conceptual guidance to that problem that can be applied countless times without ever being executed in the same way twice.

Definition of the Subject Glossary Biophilia Mankind’s innate biological connection with nature; the urge to affiliate with other forms of life. Biophilic design The process of basing decisions about the built environment on intuition or credible research – derived from either an appetency for nature or measurable biological responses, respectively – to achieve the best possible health outcomes. Biophilic design pattern A description of a pattern in nature that engenders a positive biological response in humans. Meant to inform, guide, and assist in the design process for the built environment. Built environment Places and spaces created or modified by people with particular emphasis on buildings, parks, streetscapes, and other spaces that provide the setting for human activity. Connection to place The emotional bond between person and place. Evidence-based design The process of basing decisions about the built environment on

Biophilia is humankind’s innate biological connection with nature. In the face of contemporary concerns for individual and public health and wellbeing – most typically occupational stress, cognitive performance, and mental health – design strategies that embrace qualities from nature have emerged as a legitimate means to enhance the human experience of the built environment. An underlying premise of incorporating nature into the built environment is that when people are regularly in contact with nature, personal health and wellbeing will improve in a manner and to a degree that contributes meaningfully to public health, community resilience, and environmental stewardship. An increased focus on combatting childhood health and behavioral disorders, indoor environmental quality, workplace stress, and community cohesion in recent years has provided impetus for progressively more creative preventative health strategies for building and landscape design and urban planning. As the world population continues to urbanize, design that mitigates stress, improves cognitive function and creativity, and expedites healing is

© Springer Science+Business Media, LLC, part of Springer Nature 2020 V. Loftness (ed.), Sustainable Built Environments, https://doi.org/10.1007/978-1-0716-0684-1_1034 Originally published in R. A. Meyers (ed.), Encyclopedia of Sustainability Science and Technology, © Springer Science+Business Media LLC 2018 https://doi.org/10.1007/978-1-4939-2493-6_1034-1

43

44

Biophilic Design

ever more important. Given how quickly an experience of nature can elicit a restorative response, and that US businesses squander billions of dollars each year on lost productivity due to stressrelated illnesses (Fig. 1), design that reconnects us with nature – biophilic design – is essential for providing people opportunities to live and work in healthy places and spaces with less stress and greater overall health and wellbeing. An important indirect benefit of biophilic design is the emphasis placed on identifying qualitative and quantitative science-based metrics to guide design practices and to assess efficacy after implementation, informing the dynamic and evolving relationship between user groups, building typologies, and desired health outcomes. For decades, research scientists and design practitioners have been working to define aspects of nature that most impact our satisfaction with the built environment. Yet transitioning from

research to application in a manner that effectively enhances health and wellbeing with efficacy continues to be a challenge. While there is general consensus among early adopters that biophilic design is a valued and helpful tool for approaching healthful design, there remains a significant gap between industry awareness and effective articulation of a science-based design intent. Moreover, few educational institutions have adopted curricula to introduce biophilic design as an interdisciplinary subject.

Biophilic Design, Fig. 1 Re-engaging losses from unproductive operating costs [1]. More than 90% of a company’s operating costs are linked to human resources, and financial losses due to absenteeism and presenteeism (going to work sick) account for as much as 4% of these employment costs. Commercial spaces that give occupants

access to nature to serve as a release to environmental stress offer potential savings of thousands of dollars per year in employee costs (Statistics: US Department of Labor [2], Bureau of Labor Statistics [3]; BOMA [4]; Photo: Bilyana Dimitrova; Graphic: Terrapin Bright Green)

Introduction The biophilia hypothesis [5] helps explain why crackling fires and crashing waves captivate; why a garden view can enhance creativity; why shadows and heights instill fascination and fear; and why animal companionship and strolling

Biophilic Design

through a park have restorative, healing effects. Biophilia may also help explain why some urban parks and buildings are preferred over others. Nature themes can be found in the earliest human structures: from the stylized animals characteristic of the Neolithic Göbekli Tepe; the Egyptian sphinx; the acanthus leaves adorning Greek temples; and the primitive earthen hut, to the delicate, leafy filigrees of Rococo design. Representations of animals and plants have long been used for decorative and symbolic ornamentation. Beyond representation, cultures around the world have long brought nature into homes and public spaces. Classic examples include the garden courtyards of the Alhambra in Spain, porcelain fish bowls in ancient China, the aviary in Teotihuacan (ancient Mexico City), Japanese bonsai, papyrus ponds in the homes of Egyptian nobility, the cottage garden in medieval Germany, and the elusive hanging gardens of Babylon. The consistency of natural themes in historic structures and places suggests that biophilic design is not a new phenomenon; rather, as a tool for applied science, it is the codification of history, human intuition, and neural sciences showing that connections with nature are vital to maintaining a healthful and vibrant existence as an urban species. Prior to and even after the Industrial Revolution, the vast majority of humans lived an agrarian existence, spending much of their lives among nature. As urban populations grew in the nineteenth century, reformers became increasingly concerned with health and sanitation issues such as fire hazards and dysentery. The creation of large public parks became a campaign to improve the health and reduce the stress of urban living. During that time, American landscape architect Frederick Law Olmsted argued “. . .the enjoyment of scenery employs the mind without fatigue and yet exercises it, tranquilizes it and yet enlivens it; and thus, through the influence of the mind over the body, gives the effect of refreshing rest and reinvigoration to the whole system” [6]. Artists and designers of the Victorian era, such as the influential English painter and art critic, John Ruskin, pushed back against what they saw as the dehumanizing experience of industrial cities. They

45

argued for objects and buildings that reflected the hand of the craftsman and drew from nature for inspiration. In the design of the Science Museum at Oxford, Ruskin is said to have told the masons to use the surrounding countryside for inspiration; this perspective was memorialized in the hand-carved flowers and plants adorning the museum [7]. Western attitudes toward nature shifted in the mid-nineteenth century; natural landscapes became valid art subjects, as seen in the Hudson River School in the United States and the Barbizon School in France. Going to the mountains or seashore for recreation was becoming a popular pastime; winter gardens and conservatories became requisites of wealthy homes in Europe and the United States. Hospital design of the era acknowledged the role of sunlight and a view to nature in healing; Dr. Thomas Kirkbride “. . .believed that the beautiful setting. . .restored patients to a more natural balance of the senses” [8]. Henry David Thoreau built a cabin by Walden Pond in Concord, Massachusetts, from which he wrote treatises on a simpler life connected to nature, which still resonates in the American consciousness. Art Nouveau designs of the late nineteenth century were expressly inspired by nature. Architect Victor Horta’s exuberant plant tendrils lacing through buildings in Belgium, the lush flowers that are Louis Comfort Tiffany lamps, and the explicitly organic forms of Antonio Gaudí’s buildings all remain powerful examples. In Chicago, Louis Sullivan created elaborate building façade ornamentation with leaves and cornices that represent tree branches. Sullivan’s protégé, Frank Lloyd Wright, was part of the group that launched The Prairie School. Wright abstracted prairie flowers and plants for his art glass windows and ornamentation. Like many in the Craftsman movement, Wright used the grain of wood and texture of brick and stone as decorative architectural elements. Wright also opened up residential interiors to greater spatial flow in ways that had not been done before, creating prospect views balanced with intimate refuges. His later designs sometimes included exhilarating spaces; the balcony cantilevering out over the waterfall at Fallingwater is a classic example.

46

Biophilic Design

`European Modernists stripped much ornamentation from their buildings, but, like Wright, used wood grain and the veining of stone as decorative elements. They too were equally concerned with exploring the relationship between interior and exterior spaces. Ludwig Mies van der Rohe’s Barcelona Pavilion (built 1929) pushed that concept further through the play of volumes and glass. Later, Mies’s Farnsworth House (built 1951) defined interior and exterior more literally by segregating the elements from a visual connection to nature. Le Corbusier’s Cité Radiant (unbuilt 1924) placed towers in a park surrounded by grass and trees, trying (albeit unsuccessfully) to provide city dwellers with a connection to nature. Indeed, as the International Style took root in the midtwentieth century, it populated the landscape with steel and glass buildings, which increasingly disconnected people from nature over time, particularly as building became taller. Social psychologist Eric Fromm is credited with coining the term “biophilia” in his 1964 publication The Heart of Man [9]. Biologist Edward O. Wilson later popularized the term “biophilia” with an eponymous title in 1984 [10] and The Biophilia Hypothesis in 1993 [5]. The sundry denotations – which have evolved from within the fields of biology and psychology and been adapted to the fields of neuroscience, endocrinology, architecture, and beyond – all relate back to the desire for a (re)connection with nature and natural systems. That humans should be genetically predisposed to prefer certain types of nature and natural scenery, specifically the savanna analogue, was posited by Gordon Orians and Judith Heerwagen [11, 12] and could theoretically be a contributing motivation for moving to the suburbs, with the suburban lawn being a savanna for everyone. Several theories have since been leveraged to varying degrees to set the foundation for biophilic design: • • • • • • •

Habitat and Prospect-refuge theories [13] Ecological aesthetic theory [14–16] Attention restoration theory [17] View preference matrix [17] Savanna hypothesis [11, 12] Biophilia hypothesis [5] Environmental generational amnesia [18]

With the emergence of the green building movement in the early 1990s, linkages were made between improved environmental quality and worker productivity [19]. While the financial gains due to productivity improvements were considered significant, productivity was identified as a placeholder for health and wellbeing, which were perceived to have even broader impacts. The healing power of a connection with nature was established by Roger Ulrich’s landmark study comparing recovery rates of patients with and without a view to nature [20]. An experiment at a new Herman Miller manufacturing facility, designed by William McDonough + Partners in the 1990s, was one of the first to specifically frame the mechanism for gains in productivity to connecting building occupants to nature – phylogenetic or, more familiarly, biophilic design [21]. The early twenty-first century has seen a steady growth in the characterization of the intersections of neuroscience and architecture, both in research and in practice. Popular texts, particularly Last Child in the Woods [22], Healing Spaces [8], Biophilic Design [23], The Economics of Biophilia [1], 14 Patterns of Biophilic Design [24], and The Practice of Biophilic Design [25], have brought the conversation further into the mainstream, helping the public grapple with modern society’s dependency on technology and persistent disconnect with nature. Green building stan® dards, such as the WELL Building Standard and Living Building Challenge, have begun to incorporate biophilia, predominantly for contributing to indoor environmental quality and connection to place. Most recently, biophilic design is being championed as a complementary strategy for addressing workplace stress, student performance, patient recovery, community cohesiveness, and other familiar challenges to health and overall well-being (Fig. 2).

Defining Nature Views of what constitutes natural, nature, and wilderness vary greatly, from that which can be classified as a living organism or system of organisms unaffected by anthropogenic impacts on the

Biophilic Design

Biophilic Design, Fig. 2 Thought leadership [23]. Biophilic Design was the first comprehensive publication on the theory and practice of biophilic design

environment, to all that is designed or made as an extension of the human phenotype. This broad room for interpretation adds complexity to defining parameters for expressions of biophilic design. Some articulation of what is meant by “nature” has given greater context to academics and practitioners of biophilic design. Nature is defined here as the living organisms and the nonliving components of an ecosystem. For added clarity, the distinction is made that most nature in modern society is designed, whether deliberately (for function or aesthetic), haphazardly (for navigability or access to resources), or passively (through neglect or hands-off preservation). Referring back to humankind’s proclivity for the savanna landscape, new savanna analogues are created all the time. Some designed ecosystems are bio-diverse, vibrant, and ecologically healthy, such as the high canopy forests with floral undergrowth maintained by the annual burning practices of the Ojibwe people of North America. Others are monocultures like suburban lawns and golf courses that, while beautiful savanna analogues, are not necessarily bio-diverse, ecologically healthy or resilient.

47

Transition from Theory to Application Efforts to transition from theory to application began to emerge after the publication of Building for Life: Designing and Understanding the Human-Nature Connection by Stephen Kellert in 2005 [26], and Biophilic Design: The Theory, Science and Practice of Bringing Buildings to Life in 2008 [23]. With limited awareness, market demand, or measurable return on investment, early adoption of biophilia was slow until the 2012 release of The Economics of Biophilia: Why Designing with Nature in Mind Makes Financial Sense [1], which shed light on the financial advantages of biophilic design across healthcare, education, retail, workplace, and community environments. More recently, organizations are forming to support the biophilic design movement. The Biophilic Cities Initiative has 16 member cities from around the world that share design and regulatory strategies to reconnect people with nature in urban settings. The Biophilic Design Initiative of the International Living Future Institute (ILFI) is an effort to connect designers, researchers, and others in the support of education and research in the practice of biophilic design. These groups and others are developing frameworks for identifying design-transferable patterns in nature, capturing case studies for best building practices, and promoting new peer-reviewed research as practical tools for broadening awareness, adoption, and informed application. There are several building industry perspectives that continue to promote the application of biophilic design strategies: • Occupant health and wellness is increasingly being recognized by corporations, real estate managers, and property owners as a strategy for enhancing the occupant experience and attracting and retaining tenants and employees. • Biophilic design is increasingly being recognized by the design community as a tool for addressing the growing demand for scalable health and wellness solutions for their projects. • Where competition drives innovation, biophilic design is viewed as a science-backed differentiator, providing a competitive advantage in the marketplace.

48

• As wearable technologies become more financially accessible, small-scale replicable research methods are being used to: streamline consistent metrics; expedite data collection, reporting, and design decision making and to encourage ownership by citizens of their own lifestyle and design choices. Early proponents of biophilic design literacy formulated principles for design implementation which were articulated by Kellert and Calabrese [25] as “the fundamental conditions for the effective practice of biophilic design” [25, pp. 6–7]. Principles of Biophilic Design

1. Biophilic design requires repeated and sustained engagement with nature. 2. Biophilic design focuses on human adaptations to the natural world that over evolutionary time have advanced people’s health, fitness, and wellbeing. 3. Biophilic design encourages an emotional attachment to particular settings and places. 4. Biophilic design promotes positive interactions between people and nature that encourage an expanded sense of relationship and responsibility for the human and natural communities. 5. Biophilic design encourages mutually reinforcing, interconnected, and integrated architectural solutions. Nature-Design Relationships The translation of biophilia as a hypothesis into design of the built environment was the topic of a 2004 conference and subsequent book on biophilic design [23] in which Stephen Kellert identified more than 70 different mechanisms for engendering a biophilic experience. In the same title, contributing authors William Browning and Jenifer Seal-Cramer outlined three classifications of the user experience relative to design – Nature in the Space, Natural Analogues, and Nature of the Space – providing a framework for understanding and enabling thoughtful incorporation of a rich diversity of strategies into the built environment. The 14 design patterns presented in

Biophilic Design

Browning et al. [24] utilizes using this categorization methodology [24]. Nature in the Space. Nature in the Space addresses the direct, physical, and ephemeral presence of nature in a space or place. This includes plant life, water, and animals, as well as breezes, sounds, scents, and other natural elements. Common examples include potted plants, flowerbeds, bird feeders, butterfly gardens, water features, fountains, aquariums, courtyard gardens, and green walls or vegetated roofs. The strongest Nature in the Space experiences are achieved through the creation of meaningful, direct connections with these natural elements, particularly through diversity, movement, and multisensory interactions. Natural Analogues. Natural Analogues addresses organic, nonliving, and indirect evocations of nature. Objects, materials, colors, shapes, sequences and patterns found in nature, manifest as artwork, ornamentation, furniture, décor, and textiles in the built environment. Mimicry of shells and leaves, furniture with organic shapes, and natural materials that have been processed or extensively altered (e.g., wood planks, granite tabletops) each provides an indirect connection with nature: while they are real, they are only analogous of the items in their “natural” state. The strongest Natural Analogue experiences are achieved by providing information richness in an organized and sometimes evolving manner. Nature of the Space. Nature of the Space references spatial configurations experienced in nature. This includes an individual’s innate and learned desire to be able to see beyond immediate surroundings, fascination with the unknown, and obscured views and revelatory moments. Phobiainducing spatial properties are also suggested to have a restorative benefit when a trusted element of safety is also present. The strongest Nature of the Space experiences are achieved through the creation of deliberate and engaging spatial configurations commingled with patterns of Nature in the Space and Natural Analogues. Nature-Health Relationships Much of the evidence for biophilia can be linked to research in one or more of three overarching

Biophilic Design

mind-body systems – cognitive, psychological, and physiological – that have been explored and verified to varying degrees in laboratory or field studies to help explain how human health and well-being are impacted by the physical environment. See Table 1 for nature-health relationships and biophilic design patterns. Cognitive Functionality and Performance. Cognitive functioning encompasses an individual’s mental agility and memory, and ability to think, learn, and output either logically or creatively. For instance, directed attention is required for many repetitive tasks, such as routine paperwork, reading, and performing calculations or analysis, as well as for operating in highly stimulating environments, as when crossing busy streets. Directed attention is energy intensive and, over time, can result in mental fatigue and depleted cognitive resources [23, 26]. Strong or routine connections with nature can provide opportunities for mental restoration, providing time when higher cognitive functions are relieved, allowing an individual’s capacity for performing focused tasks to be greater than another’s with fatigued cognitive resources. Psychological Health and Wellbeing. Psychological responses encompass an individual’s adaptability, alertness, attention, concentration, as well as emotion and mood. Psychological responses to nature impact restoration and stress management. Empirical studies have reported that experiences of natural environments provide greater emotional restoration, with lower instances of tension, anxiety, anger, fatigue, confusion, and total mood disturbance than urban environments with limited characteristics of nature [30, 32, 49, 83]. Psychological responses can be learned or hereditary, with past experiences, cultural constructs, and social norms playing a significant role in the psychological response mechanism. Physiological Health and Wellbeing. Physiological responses encompass aural, musculoskeletal, respiratory, circadian systems, and overall physical comfort. Physiological responses triggered by connections with nature include relaxation of muscles, as well as lowering of diastolic blood pressure and stress hormone (i.e.,

49

cortisol) levels in the blood stream [31]. Shortterm stress that increases heart rate and stress hormone levels, such as from encountering an unknown but complex and information-rich space, or looking over a banister to eight stories below, are suggested to be beneficial to regulating physiological health [84]. The physiological system needs to be tested regularly, but only enough for the body to remain resilient and adaptive. Physiological responses to environmental stressors can be buffered through design, allowing for the restoration of bodily resources before system damage occurs [85].

Framework Methodology The work of Christopher Alexander, Judith H. Heerwagen, Rachel and Stephen Kaplan, Stephen R. Kellert, Roger Ulrich, and others has been instrumental in framing the conversation on biophilic design for health and wellbeing. In the two decades since Kellert and Wilson [5] published The Biophilia Hypothesis [5], the body of evidence supporting biophilia has expanded considerably. Hundreds of publications on biophilic responses have since been mined to uncover unique patterns useful to designers of the built environment. Synthesis and translation of this research into a digestible list of design patterns, spearheaded by William D. Browning and Catherine O. Ryan, provided comprehensive yet flexible guidance for site-appropriate implementation. This body of work is known as the 14 patterns of biophilic design [24]: Nature in the Space encompasses seven biophilic design patterns: 1. Visual Connection with nature. A view to elements of nature, living systems, and natural processes. 2. Nonvisual connection with nature. Auditory, haptic, olfactory, or gustatory stimuli that engender a deliberate and positive reference to nature, living systems, or natural processes. 3. Nonrhythmic sensory stimuli. Stochastic and ephemeral connections with nature that may be

50

Biophilic Design

Biophilic Design, Table 1 Biophilic design patterns and biological responses [24, p. 12] 14 patterns Nature in the space

* * * * * *

Stress reduction Lowered blood pressure and heart rate [26–28] Reduced systolic blood pressure and stress hormones [31–34]

Cognitive performance Improved mental engagement/ attentiveness [29] Positively impacted cognitive performance [35, 36]

Nonrhythmic sensory stimuli

* *

Thermal and airflow variability

* *

Positively impacted heart rate, systolic blood pressure, and sympathetic nervous system activity [34, 42–45] Positively impacted comfort, well-being and productivity [47, 48, 132]

Observed and quantified behavioral measures of attention and exploration [46] Positively impacted concentration [17, 32, 49]

Presence of water

* *

Reduced stress, increased feelings of tranquility, lower heart rate, and blood pressure [29, 54, 55]

Improved concentration and memory restoration [29, 54] Enhanced perception and psychological responsiveness [54, 56]

Dynamic and diffuse light

* *

Positively impacted circadian system functioning [62, 63] Increased visual comfort [64; 128]

Visual connection with nature Nonvisual connection with nature

Connection with natural systems Natural analogues

Nature of the space

Biomorphic forms and patterns Material connection with nature

*

Complexity and order

* *

Prospect

* * * * * *

Refuge

Emotion, mood and preference Positively impacted attitude and overall happiness [30] Perceived improvements in mental health and tranquility [37–41]

Improved perception of temporal and spatial pleasure (alliesthesia) [50–53]; [129, 130] Observed preferences and positive emotional responses [29, 30, 46, 57–61]

Enhanced positive health responses; Shifted perception of environment [23] Observed view preference [65] Decreased diastolic blood pressure [66] Improved creative performance [67] Positively impacted perceptual and physiological stress responses [65, 68, 69, 131] Reduced stress [72]

Improved comfort [66, 80]

Observed view preference [68–71]

Reduced boredom, irritation, fatigue [73]

Improved comfort and perceived safety [74–76]

Improved concentration, attention and perception of safety [72, 74, 75, 77] (continued)

Biophilic Design

51

Biophilic Design, Table 1 (continued) 14 patterns Mystery Risk/Peril

* * * *

Stress reduction

Cognitive performance

Emotion, mood and preference Induced strong pleasure response [78–80] Resulted in strong dopamine or pleasure responses [81, 82, 133]

© 2014 Terrapin Bright Green/14 Patterns of Biophitic Design: Improving Health and Wellbeing in the Built Environment

4.

5.

6.

7.

analyzed statistically but may not be predicted precisely. Thermal and airflow variability. Subtle changes in air temperature, relative humidity, airflow across the skin, and surface temperatures that mimic natural environments. Presence of water. A condition that enhances the experience of a place through seeing, hearing or touching water. Dynamic and diffuse light. Leverages varying intensities of light and shadow that change over time to create conditions that occur in nature. Connection with natural systems. Awareness of natural processes, especially seasonal and temporal changes characteristic of a healthy ecosystem (Fig. 3).

Natural Analogues encompasses three patterns of biophilic design: 8. Biomorphic Forms and Patterns. Symbolic references to contoured, patterned, textured or numerical arrangements that persist in nature. 9. Material connection with nature. Materials and elements from nature that, through minimal processing, reflect the local ecology or geology and create a distinct sense of place. 10. Complexity and order. Rich sensory information that adheres to a spatial hierarchy similar to those encountered in nature (Fig. 4). Nature of the Space encompasses four biophilic design patterns:

11. Prospect. An unimpeded view over a distance, for surveillance and planning. 12. Refuge. A place for withdrawal from environmental conditions or the main flow of activity, in which the individual is protected from behind and overhead. 13. Mystery. The promise of more information, achieved through partially obscured views or other sensory devices that entice the individual to travel deeper into the environment. 14. Risk/Peril. An identifiable threat coupled with a reliable safeguard (Fig. 5). From extensive interdisciplinary research, key patterns in nature – those known, suggested, or theorized to mitigate common stressors or enhance desirable qualities – were identified in relation to design patterns for application in the built environment. These biophilic design patterns have, in the words of Wilson, been “teased apart and analyzed individually” to reveal the emotional affiliations Wilson spoke of, as well as other psycho-physiological and cognitive relationships with the built environment. The descriptive term “pattern” is being used for three reasons: 1. To propose a clear and standardized terminology for biophilic design 2. To avoid confusion with multiple terms (metric, attribute, condition, characteristic, typology, etc.) that have been used to explain biophilia and biophilic design 3. To maximize accessibility across disciplines by upholding a familiar language

52

Biophilic Design, Fig. 3 Nature in the space. Traditional Japanese and Chinese gardens are very successful models for bringing nature into a space with diverse plantings,

Biophilic Design

water features, animals, varying depths of view, and hints of human activity (Credit: Oskar Vertetics, common use)

Biophilic Design, Fig. 4 Natural analogues. Metropol Parasol, by J Mayer H architects, is an undulating, biomorphic timber structure amidst a dense urban area of Seville, Spain (Credit: Oliver Gobet, common use)

Biophilic Design

53

Biophilic Design, Fig. 5 Nature of the space. The boundary between indoors and outdoors is blurred from the Farnsworth House by Ludwig Mies van der Rohe. While the structure itself is not particularly biophilic, the

experience of the spatial qualities in relation to the surrounding environment affords the occupant a sense of prospect and refuge (Credit: Marc Teer/flickr)

The use of spatial patterns is inspired by the precedents of A Pattern Language [86], Designing with People in Mind [87], and Patterns of Home [88], as well as lectures and compilations on patterns, form, language and complexity [89, 90]. Christopher Alexander brings clarity to this intent with his explanation that patterns “. . .describe a problem which occurs over and over again in our environment, and then describes the core of the solution to that problem, in such a way that you can use this solution a million times over, without ever doing it the same way twice” [86]. Alexander’s work built on the tradition of pattern books used by designers and builders from the eighteenth century onward, but his work focused on the psychological benefits of patterns and included descriptions of the three dimensional spatial experience, rather than the aesthetic focus of previous pattern books. With Terrapin Bright Green’s fourteen patterns of biophilic design, the focus is on psychological, physiological, and cognitive benefits that address universal issues of

human health and wellbeing (e.g., stress, visual acuity, hormone balance, perceived comfort, creativity), rather than those unique to programbased or sector-specific space types (e.g., health care facility waiting rooms, elementary school classrooms, or storefront pedestrian promenades). Each design pattern was assessed for overall potential impact and the strength of the research on which a pattern is built. The Biophilic Design Patterns and Biological Responses matrix (Table 1) was one of the first visual tools to communicate relationships between the research, health outcomes, and design qualities of the built environment. Table 1 illustrates the functions of each of the 14 Patterns in supporting stress reduction, cognitive performance, emotion and mood enhancement, and support of the human body. Patterns that are supported by more rigorous empirical data are marked with up to three asterisks (***), indicating that the quantity and quality of available peer-reviewed evidence is robust and the potential

54

for impact is great. No asterisk indicates that there is minimal research to support the biological relationship between health and design, but the anecdotal information is compelling and adequate for hypothesizing its potential impact and importance as a unique pattern.

Design Patterns and Implementation While informed by science, biophilic design patterns are not formulas; they are meant to inform, guide, and assist in the design process and should be thought of as another tool in the designer’s toolkit. The purpose of defining these patterns is to articulate connections between aspects of the built and natural environments and how people react to and benefit from them. Just as combinations of culture, demographics, health baselines, and characteristics of the built environment can impact the experience of space differently, so too can each design pattern. A suitable solution results from understanding local conditions and one space’s relationship to another, and responding appropriately with a combination of design interventions to suit the unique needs of a space and its intended user group and programs. The field of biophilic design is constantly evolving. As Nikos Salingaros explained, new disciplines such as biophilic design must “abstract its patterns as they appear. . . building its own foundation and logical skeleton, upon which future growth can be supported” [90]. As new evidence comes to bear, some patterns will be championed over others and new patterns may emerge. Pattern 1. Visual Connection with Nature A Visual Connection with Nature is a view to elements of nature, living systems, and natural processes. A biophilic experience of a good Visual Connection with Nature feels whole, commands attention, and can be stimulating or calming, conveying a sense of time, weather, and other living things. Roots of the Pattern. The Visual Connection with Nature pattern has evolved from both research on visual preferences and responses to views to nature that include reduced stress, more positive

Biophilic Design

emotional functioning, and improved concentration and recovery rates. Stress recovery from visual connections with nature has reportedly been realized through lowered blood pressure and heart rate; reduced sadness, anger, and aggression; improved mental engagement/attentiveness; and improved attitude and overall happiness [26–30]. There is evidence showing that stress reduction can be attributed to both experiencing real nature and seeing images of nature [32, 42, 72, 91, 92] and that natural environments are generally preferred over built environments [93–95]. Visual preference research indicates that the preferred view is looking down a slope to a scene that includes copses of shade trees, flowering plants, calm nonthreatening animals, indications of human habitation, and bodies of clean water [11]. This is often difficult to achieve in the built environment, particularly in already dense urban settings, though the psychological benefits of nature are suggested to increase with higher levels of biodiversity and not with an increase in natural vegetative area [96]. Positive impact on mood and self-esteem has also been shown to occur most significantly in the first 5 min of experiencing nature, such as through exercise within a green space [30]. Viewing nature for 10 min prior to experiencing a mental stressor has shown to positively stimulate heart rate variability and parasympathetic activity (i.e., regulation of internal organs and glands that support digestion and other activities that occur when the body is at rest) [27], while viewing a forest scene for 20 min after a mental stressor has shown to return cerebral blood flow and brain activity to a relaxed state [28]. Viewing scenes of nature stimulates a larger portion of the visual cortex than nonnature scenes, which triggers more pleasure receptors in the brain, leading to prolonged interest and faster stress recovery – within as little of 40 s [97]. For example, heart rate recovery from low-level stress, such as from working in an office environment, has shown to occur 1.6 times faster when the space has a glass window with a nature view, rather than a high-quality simulation (i.e., plasma video) of the same nature view, or no view at all [42]. Additionally, repeated viewing of real nature, unlike nonnature, does not significantly diminish the viewer’s level of interest over time [29].

Biophilic Design

Design Objectives and Guidance. The objective of the Visual Connection with Nature pattern is to provide an environment that helps the individual change visual focus to relax the eye muscles and temper cognitive fatigue. The effect of an intervention will improve as the quality of a view and the amount of visible biodiversity each increases. A view to nature through a glass window provides a benefit over a digital screen (e.g., video/ plasma tv) of the same view, particularly because there is no parallax shift for people as they move toward or around a video screen [42]. This may change as three-dimensional videography advances. Both simulated and constructed nature is measurably better at engendering stress reduction than having no visual connection at all. Design guidance that can be translated from research to help establish a strong visual connection with nature includes: • Prioritization of real nature over simulated nature; and simulated nature over no nature. • Prioritization of biodiversity over acreage, area or quantity. • Accessibility to exercise opportunities in green space. • Design to support a visual connection that can be experienced for at least 5–20 min/day. • Design spatial layouts and furnishings to uphold desired view lines and avoid impeding the visual access when in a seated position. • Infusion of restorative instances of nature in small spaces, particularly for spaces where real estate (floor/ground area, wall space) is limited. • Inclusion of digital access to nature in spaces where real nature or views to the outdoors cannot be provided (e.g., healthcare radiation units). The benefits of viewing real nature may be attenuated when experienced through a digital medium; however, access to digitized or static images of nature is better than no visual connection (Fig. 6). Pattern 2. Nonvisual Connection with Nature Nonvisual Connection with Nature is the auditory, haptic, olfactory, or gustatory stimuli that engender a deliberate and positive reference to nature, living systems, or natural processes. A multisensory biophilic experience feels fresh and well balanced;

55

the ambient conditions are perceived as complex and variable but at the same time familiar and comfortable, with sounds, aromas, and textures reminiscent of being outdoors in nature. Roots of the Pattern. The Nonvisual Connection with Nature pattern has evolved from data on reductions in systolic blood pressure and stress hormones [31–33], impact of sound and vibration on cognitive performance [35, 36], and perceived improvements in mental health and tranquility as a result of nonvisual sensory interactions with nonthreatening nature [37–41]. Each sensory system has an independent body of research to support it; a brief overview provides some context. Auditory. Research shows that exposure to nature sounds, when compared to urban or office noise, accelerates physiological and psychological restoration up to 37% faster after a psychological stressor [54] and reduces cognitive fatigue and helps motivation. Participants of one study who either listened to river sounds or saw a nature movie with river sounds during a posttask restoration period reported having more energy and greater motivation after the restoration period compared to participants who only listened to office noise or silence [38]. In addition, viewing the nature movie with river sounds during the restoration period had a more positive effect than only listening to river sounds alone. Ocean waves and vehicle traffic can have a very similar sound pattern. In an experiment using a synthesized sound that replicated the waves and traffic sound pattern, researchers observed that participants processed the synthesized sound in different portions of the brain depending on whether they were also watching a video of either waves or vehicle traffic [56]. Participants considered the sound to be pleasurable when viewing the video of waves, but not when viewing the video of traffic. This study suggests a strong connection between visual and auditory sensory systems and psychological well-being. Olfactory. Our olfactory system processes scent directly in the brain, which can trigger very powerful memories. Traditional practices have long used plant oils to calm or energize people. Studies have also shown that olfactory exposure to herbs and phytoncides (essential oils from

56

Biophilic Design

Biophilic Design, Fig. 6 Visual connection with nature. An example of a designed environment with an excellent Visual Connection with Nature is the birch tree and moss garden in the New York Times Building in New York City – a carved out space in the middle of the building

which everyone passes as they enter or leave the building. Adjacent to a restaurant and the main conference rooms, the birch garden is an oasis of calm in the hustle of Times Square. (Source: John Zacherle)

trees) has a positive effect on the healing process and human immune function, respectively [37, 40]. Haptic. Pet therapy, where companionship and the act of activities petting and feeling the fur of domesticated animals, is known to have profound calming effects on patients; gardening and horticulture have shown to engender environmental stewardship among children, reduce self-reported fatigue while maintaining joint flexibility among adults [98], and reduce perception of pain among senior populations with arthritis. The act of touching real plant life, versus synthetic plants, has also shown to induce relaxation through a change in cerebral blood flow rates [99]. These examples give reason to believe that the experience of touching other elements in nature, such as water or raw materials, may result in similar health outcomes. Gustatory. Tasting is yet another way of experiencing nature and learning about the environment. While adults are often curious or fearful of edible plants and herbs, consider the familiar habit of infants and toddlers putting found objects in their mouths – they are seeking information.

Design Objectives and Guidance. The objective of the Nonvisual Connection with Nature pattern is to provide an environment that uses sound, scent, touch, and possibly even taste to engage the individual in a manner that helps reduce stress and improve perceived physical and mental health. These senses can be experienced separately, although the experience is intensified and the health effect is compounded if multiple senses are consistently engaged together. Design guidance translated from research to help establish a strong nonvisual connection with nature and maximize potential positive health responses: • Prioritization of nature sounds over urban sounds. • Design for nonvisual connections that can be easily accessed from one or multiple locations, and in such a way that allows daily engagement for 5–20 min at a time. • Integration of nonvisual connections with other aspects of the design program.

Biophilic Design

Biophilic Design, Fig. 7 Nonvisual connection with nature. Calat Alhambra in Granada, Spain, is an exquisite example of the 14 Patterns. While some patterns are more evident in some spaces than others, Nonvisual Connections with Nature are experienced throughout. The integration of water and natural ventilation with the architecture is central to the nonvisual experience, supporting a seamless connection between indoor and outdoor spaces and between the building and the surrounding natural landscape. Solar heat penetrates at distinct locations, the whispering gallery resonates sounds of nature and people, and gardens of rosemary, myrtles, and other fragrant plants surround the premises. The extensive use of water fountains creates a microclimate – the space both sounds and feels cooler – while stone floors and handrails with water channels cool the feet and hands through conductance (Credit: Sharon Mollerus/flickr (creative commons))

• A single intervention that can be experienced in multiple ways. • Design for visual and nonvisual connections to be experienced simultaneously (Fig. 7). Pattern 3. Nonrhythmic Sensory Stimuli Nonrhythmic Sensory Stimuli are stochastic and ephemeral connections with nature that may be analyzed statistically but may not be predicted

57

precisely. A biophilic experience of NonRhythmic Sensory Stimuli is a brief but welcome distraction, an ephemeral intrigue that may be perceived as stimulating and energizing. Roots of the Pattern. The Nonrhythmic Sensory Stimuli pattern has evolved from research around looking behavior (particularly periphery vision movement reflexes); eye lens focal relaxation patterns; heart rate, systolic blood pressure, and sympathetic nervous system activity [34, 42–45]; and observed and quantified behavioral measures of attention and exploration [46]. The movement of living things and mechanical objects are each processed in different parts of our brain [44], whereby natural movement is generally perceived as positive, and mechanical movement as neutral or even negative. As a result, the repeating rhythmic motion of a pendulum will only hold one’s attention briefly, the constant repetitive ticking of a clock may come to be ignored over time, and an ever-present scent may lose its mystique with long-term exposure, whereas the stochastic movement of a butterfly will capture one’s attention each time, even for a brief moment, for recurring physiological benefits. Studies of the human response to stochastic movement of objects in nature and momentary exposure to natural sounds and scents have shown to support physiological restoration. For instance, when sitting and staring at a computer screen or doing any task with a short visual focus, the eye’s lens becomes rounded with the contracting of the eye muscles. When these muscles stay contracted for an extended period, i.e., more than 20 min at a time, fatigue can occur, manifesting as eye strain, headaches, and physical discomfort. A periodic, yet brief visual or auditory distraction that causes one to look up (for >20 s) and to a distance (of >20 ft) allows for short mental breaks during which the muscles relax and the lenses flatten. Design Objectives and Guidance. The objective of the Nonrhythmic Sensory Stimuli pattern is to encourage the use of natural sensory stimuli that unobtrusively attract attention, allowing individuals’ capacity for focused tasks to be replenished from mental fatigue and physiological stressors. This can be achieved by designing for momentary exposure to the stochastic or unpredictable movement, particularly for

58

periphery vision or the periodic experience of scents or sounds. When immersed in nature, instances of nonrhythmic stimuli are continually experienced: birds chirping, leaves rustling, and the faint scent of eucalyptus in the air. The built environment has evolved into a deliberately predictable realm. Even some highly manicured gardens and certainly interior vegetation lack the qualities needed to support nonrhythmic sensory stimuli. Design guidance for establishing accessible and effective nonrhythmic stimuli: • As a general guideline, nonrhythmic sensory experiences should occur approximately every 20 min for about 20 s and for visual stimuli, from a distance of more than 20 ft away. • Collaboration with the mechanical engineer or facilities team may be necessary early in projects that use mechanized simulation (rather than natural occurrence) of natural stimuli.

Biophilic Design, Fig. 8 Nonrhythmic sensory stimuli. The Dockside Green community on Vancouver Island, Victoria, BC Canada, is a great example of nonrhythmic stimuli. The implementation of habitat restoration and rainwater management has led to ephemeral experiences

Biophilic Design

• Collaboration with landscape or horticulture planning will help ensure an appropriate and effective application of nonrhythmic stimuli. For instance, the selection of plant species for window boxes that will attract bees, butterflies, and other pollinators may be a more practical application for some projects than maintaining a honeybee apiary or butterfly sanctuary. • The seasonality of some naturally occurring stimuli may necessitate a variety of interventions to help ensure effective nonrhythmic sensory experiences can occur at any given time. • Orientation of stimuli relative to the occupant is in the peripheral view, from where it will be perceived much faster than from straight ahead (Fig. 8).

Pattern 4. Thermal and Airflow Variability Thermal and Airflow Variability can be characterized as subtle changes in air temperature, relative

of swaying grasses, falling water and the buzz of passing insects and animals that are visible from walkways, porches, and windows around the community (Credit: jayscratch/flickr)

Biophilic Design

humidity, airflow across the skin, and surface temperatures that mimic natural environments. A biophilic experience of Thermal and Airflow Variability feels refreshing, active, alive, invigorating, and comfortable. The space provides a feeling of both flexibility and a sense of control. Roots of the Pattern. The Thermal and Airflow Variability pattern has evolved from: research measuring the effects of natural ventilation, its resulting thermal variability, and worker comfort, well-being, and productivity [47, 48]; physiology and perception of temporal and spatial alliesthesia (pleasure) [50–53]; Attention Restoration Theory and the impact of nature in motion on concentration [32, 49]; and generally speaking, a growing discontent with the conventional approach to thermal design which focuses on trying to achieve a narrow target area of temperature, humidity, and air flow while minimizing variability [100]. Evidence has shown that people like moderate levels of sensory variability in the environment, including variation in light, sound, and temperatures and that an environment devoid of sensory stimulation and variability can lead to boredom and passivity. Early studies in alliesthesia indicate that pleasant thermal sensations are better perceived when one’s initial body state is warm or cold, not neutral [101], which corroborates more recent studies reporting that a temporary overcooling of a small portion of the body when hot, or over-heating when cold, even without impacting the body’s core temperature, is perceived as highly comfortable [51]. According to Attention Restoration Theory, elements of “soft fascination” such as light breezes or other natural movements can improve concentration [102]. Other research indicates that a variety of thermal conditions within a classroom can lead to better student performance; and that changes in ventilation velocity can have a positive impact on comfort, with no negative impact on cognitive function, while also offering the possibility of some increase in the ability to access short term memory. Design Objectives and Guidance. The objective of the Thermal and Airflow Variability pattern is to provide an environment that allows users to

59

experience the sensory elements of airflow variability and thermal variability. The intent is also for the user to be able to control thermal conditions, either by using individual controls or allowing occupants access to variable ambient conditions within a space. In contrast, conventional thermal design tries to achieve a narrow target area of temperature, humidity, and airflow, while minimizing variability – the goal being to maintain conditions within the “comfort envelope” set in the USA by the American Society of Heating, Refrigerating and Air-Conditioning Engineers (ASHRAE). When the entire space meets this goal, laboratory-based predictive models assert that 80% of the occupants would be satisfied at any given time – traditionally an acceptable outcome industry-wide. However, allowing users to easily adapt and modify the thermal conditions of their environment within two degrees Celsius above and below the conventional parameters for thermal comfort has been shown to improve perceived thermal comfort [103]. An alternative approach is to provide combinations of ambient and surface temperatures, humidity, and airflow, similar to the variability experienced outdoors during comfortable periods, while also providing some form of personal control (e.g., manual, digital, or physical relocation) over those conditions. Providing variable conductance materials, seating options with differing levels of solar heat gain (indoors and outdoors) or proximity to operable windows – as welcome as catching a cooling breeze on a sunny day or leaning one’s back on a warm rock on a cool day – could improve the overall satisfaction of a space. Since thermal comfort is inherently subjective and strongly varies between people, it is important to give a degree of control to individuals, which can be incorporated architecturally (e.g., access to operable windows or shades) and/or mechanically (e.g., access to localized and energy-efficient fans or heaters, and thermostat controls). When an individual experiences thermal discomfort, he or she will likely take action to adapt (e.g., put on a sweater; open a window; move to a different seat; or with little personal control, submit a complaint). Sometimes these adaptive actions are simply in

60

Biophilic Design

response to dynamic changes in personal preference to create an enhanced thermal experience. Design guidance translated from scientific research to help create quality thermal and airflow variability includes: • Incorporation of airflow and thermal variability into materials, daylighting, mechanical ventilation, and/or fenestration over space and time. • Recognition that thermal comfort is a vital bridging component between biophilic design and sustainable design, especially in the face of climate change and rising energy costs. When Thermal and Airflow Variability is implemented in a way that broadens people’s perception of thermal comfort, the potential increases for reduced energy demands for air conditioning and heating. • Integration of user controls and features (e.g., window treatments, varied seating options, operable windows) that allow occupants to easily adapt and modify the thermal conditions of their environment. • Coordination of design strategies among the project team (e.g., architect, lighting designer, and mechanical and electrical engineers) as early as the schematic design process (Fig. 9). Pattern 5. Presence of Water Presence of Water is a condition that enhances the experience of a place through the seeing, hearing, or touching of water. The biophilic experience of the Presence of Water is compelling and captivating. Fluidity, sound, lighting, proximity, and accessibility each contributes to whether a space is stimulating, calming, or both. Roots of the Pattern. The Presence of Water pattern has evolved from: research on the visual preference for and positive emotional responses to environments containing water elements [29, 30, 46, 57–60]; reduced stress, increased feelings of tranquility, lower heart rate and blood pressure, and recovered skin conductance from exposure to water features [54, 55]; improved concentration and memory restoration induced by complex, naturally fluctuating visual stimuli; and enhanced

Biophilic Design, Fig. 9 Thermal and airflow variability. Singapore’s Khoo Teck Puat Hospital by RMJM Architects is an excellent example of Thermal and Airflow Variability. The passive design of the hospital draws fresh air in from the exterior courtyards; the cool air helps maintain thermal comfort, while patients also have operable windows in their rooms, allowing for greater personal control. The façade and internal layouts are designed to enhance daylight and light/shade variability while reducing glare. Connecting, elevated exterior walkways also provide access to breezes, shade, and solar heat as desired (Credit: William D. Browning for Terrapin Bright Green)

perception and psychological and physiological responsiveness when multiple senses are stimulated simultaneously [29, 54–56]. Visual preference research indicates that a preferred view contains bodies of clean (i.e., unpolluted) water. Research has also shown that landscapes with water elicit a higher restorative response and generally have a greater preference among populations in comparison to landscapes without water. Supporting evidence has suggested that natural scenes without water and urban scenes with water elements elicit primarily equal benefits [38, 57, 58]. Research on response to activities conducted in green spaces has shown that the presence of water

Biophilic Design

prompts greater improvements in both self-esteem and mood than activities conducted in green environments without the presence of water [30]. Auditory access and perceived or potential tactile access to water also reportedly reduce stress [54, 55]. Design Objectives and Guidance. The objective of the Presence of Water pattern is to capitalize on the multisensory attributes of water to enhance the experience of a place in a manner that is soothing, prompts contemplation, enhances mood, and provides restoration from cognitive fatigue. Repeated experiences of water do not significantly diminish the level of interest over time [29], so one small water feature may be adequate. Taking advantage of the sounds created by smallscale running water, and the capacity to touch it, will amplify the desired health response with a multisensory experience. Vistas to large bodies of water or physical access to natural or designed water bodies can also have the health response so long as they are perceived as “clean” or unpolluted. Images of nature that include aquatic elements are more likely to help reduce blood pressure and heart rate than similar imagery without aquatic elements. Design guidance for optimizing the impacts of a presence of water, include: • Prioritization of a multisensory water experience to achieve the most beneficial outcome. • Prioritization of naturally fluctuating water movement over predictable movement or stagnancy. • Consideration of volume and turbulence for a water feature, as well as proximity to occupants, may differ depending on how humidity levels and acoustics are prioritized and controlled for in a given space type. • Consideration for climate conditions in rationale for incorporating water feature. Water features are often though not intrinsically water and energy intensive. Large water features implemented in a water sensitive region could be perceived as an unnecessary and extravagant use of resources, whereas passive features that are seasonal by design may be treated as an

61

Biophilic Design, Fig. 10 Presence of water. Water forms can serve as powerful mechanisms for stress reduction, placemaking, and community cohesion and resiliency, especially when people are able to engage in a multisensory experience. The interactive water pavilion in Hannoversche Munden, Germany, directly sought to tackle community challenges in a previously underutilized plaza devoid of a connection with nature (Credit: Herbert Dreiseitl)

educational opportunity for connecting people to the natural cycles of the local ecosystem (see Connection with Natural Systems, Pattern 7, Fig. 10). Pattern 6. Dynamic and Diffuse Light Dynamic and Diffuse Light leverages varying intensities of light and shadow that change over time to create conditions that occur in nature. A biophilic experience of Dynamic and Diffuse Light conveys expressions of time and movement to evoke feelings of drama and intrigue, buffered with a sense of calm.

62

Roots of the Pattern. Lighting design has long been used to set the mood for a space, and different lighting conditions elicit differing psychological responses. The impact of daylight on performance, mood, and well-being has been studied for many years, in a variety of environments, and in the overlapping fields of science and design, light has been extensively studied and written about. Early research showed that productivity is higher in well-daylighted work places, sales are higher in daylit stores, and that children performed better in daylighted classrooms with views. This research was focused on lighting and task performance and less on human physiological and psychological response. Yet daylighting has been reported to induce more positive moods and significantly less dental decay among students attending schools with quality daylighting, as compared to students attending schools with predictable electric lighting conditions [104]. Recent research has focused more heavily on illuminance fluctuation and visual comfort [64], human factors and perception of light, and impacts of lighting on the circadian system functioning [62, 63]. Sunlight changes color from yellow in the morning, to blue at midday, and red in the afternoon/evening. The human body responds to this daylight color transition in body temperature, heart rate, and circadian functioning. The higher content of blue light (similar to early morning light) produces serotonin, whereas an absence of blue light (evening light) produces melatonin. The balance of serotonin and melatonin can be linked to sleep quality, mood, alertness, depression, breast cancer, and other health conditions. Design Objectives and Guidance. The objective of the Dynamic and Diffuse Light pattern is twofold: to provide users with lighting options that stimulate the eye and hold attention in a manner that engenders a positive psychological or physiological response and to help maintain circadian system functioning. The goal should not be to create uniform distribution of light through a (boring) space, nor should there be extreme differences (i.e., glare discomfort). The human eye and the processing

Biophilic Design

of light and images within the brain are adaptable over a broad range of conditions, although there are limitations. For example, when the lighting difference between adjoining sources or surfaces has a brightness or luminance ratio of greater than 40:1, glare may occur, between task and immediate surroundings should not exceed 10:1. So while dramatic lighting differences may be great for some religious, socialization, and circulation spaces, they are not a good idea near work surfaces. Diffuse lighting on vertical and ceiling surfaces provides a calm backdrop to the visual scene. Accent lighting and other layering of light sources creates interest and depth, while task or personalized lighting provides localized flexibility in intensity and direction. These layers help create a pleasing visual environment. Just as variations in lighted surfaces are important for interpreting surfaces, conducting a variety of tasks, and safe navigation, circadian lighting is important for supporting biological health. Leveraging opportunities for illuminance fluctuation, light distribution, and light color variability that stimulate the human eye without causing discomfort will improve the quality of the user experience. In addition to the variability of light levels, the movement of light and shadows along a surface can provide positive attention, for example, the dappled light under the canopy of an aspen tree or the reflections of rippling water on a wall. These patterns tend to be fractals, and the brain is attuned to moving fractals (see Complexity and Order, Pattern 10). Design guidance for establishing a balance between dynamic and diffused lighting conditions includes: • Adoption of a layered lighting strategy to help facilitate easy interpretation of surfaces, flexibility of task spaces, and safe navigation between indoor and outdoor spaces. • Consideration of how dynamic lighting will impact intended programming within a space. Highly dynamic lighting conditions, such as with sustained movement, changing colors, direct sunlight penetration, and high contrasts,

Biophilic Design

63

Biophilic Design, Fig. 11 Dynamic and diffuse light. The Levine Residence on the outskirts of Los Angeles, California, experiences daylight from multiple orientations and shadow casting that changes over the course of the day.

Paired with multiple seating options, the design provides a dynamic lighting solution for a home living environment (Credit: Jeremy Levine Design)

may not be appropriate for spaces where activities requiring focused attention are performed. • Incorporation of circadian lighting design as critical feature of spaces occupied for extended periods of time with little to no natural daylight (Fig. 11).

what’s being seen and experienced. This pattern has a strong temporal element, which can be expressed culturally, as in the Japanese love of the ephemerality of cherry blossoms. Design Objectives and Guidance. The objective of the Connection with Natural Systems pattern is to heighten both awareness of natural properties and environmental stewardship of the ecosystems within which those properties prevail. The strategy for working with the pattern may be as simple as identifying key content in a view to nature (e.g., deciduous trees in the back yard or blossoming orchids on the window sill), or it may be a more complex integration of systems, such as by making evident the relationship between building occupant behavior and rainwater infrastructure (e.g., raingardens, bioswales, storm sewers), by regulating domestic activities (e.g., showering, laundry) during rain events. In either case, the temporal component is usually the key factor in pattern recognition and the triggering of a deeper awareness of a functioning ecosystem. Design guidance and opportunities that may help create quality connections with natural systems include:

Pattern 7. Connection with Natural Systems Connection with Natural Systems is the awareness of natural processes, especially seasonal and temporal changes characteristic of a healthy ecosystem. The biophilic experience evokes a relationship to a greater whole, making one aware of seasonality and the cycles of life. The experience is often relaxing, nostalgic, profound or enlightening, and frequently anticipated. Roots of the Pattern. There is limited scientific documentation of the health impacts associated with access to natural systems; however, much like Presence of Water (Pattern 5), this pattern is suspected to enhance positive health responses. In Biophilic Design [23], Kellert frames this as “Natural Patterns and Processes,” whereby seeing and understanding the processes of nature can create a perceptual shift in

64

• Integration of rainwater capture and treatment into the landscape design that responds to rain events. • Provision of visual access to existing natural systems with sensitivity to variations in the season, climate or time of day. Incorporation of responsive design tactics (e.g., use of materials that change form or expand function with exposure to solar heat gain, wind, rain/moisture, or shading), structures (e.g., steps wells), and land formations (e.g., bioswales, arroyos, dunes) will sometimes be necessary to achieve the desired level of awareness (Fig. 12). Pattern 8. Biomorphic Forms and Patterns Biomorphic Forms and Patterns are symbolic references to contoured, patterned, textured, or numerical arrangements that persist in nature. The biophilic experience feels interesting and comfortable, possibly captivating, contemplative, or even absorptive. Roots of the Pattern. Biomorphic Forms and Patterns has evolved from research on view preferences [65] reduced stress due to induced shift in focus and enhanced concentration. The science behind why humans have a visual preference for organic and biomorphic forms is not yet formulated. While the brain knows that biomorphic forms and patterns are not living things, they are perceived as symbolic representations of life. Right angles and straight lines are not naturally occurring phenomena. The Golden Angle, which measures approximately 137.5 , is the angle between successive florets in some flowers, while curves and angles of 120 are frequently exhibited in other elements of nature. The Fibonacci series (0, 1, 1, 2, 3, 5, 8, 13, 21, 34. . .) is a numeric sequence that occurs in many living things, plants especially. Phyllotaxy, or the spacing of plant leaves, branches and flower petals (so that new growth does not block the sun or rain from older growth) often follows in the Fibonacci series. Related to the Fibonacci series is the Golden Mean (or Golden Section), a ratio of 1:1.618 that surfaces time and again among living forms that grow and unfold in steps or rotations, such as with the arrangement of seeds in sunflowers or the spiral of seashells.

Biophilic Design

Biomorphic forms and patterns have been artistically expressed for millennia, from adorning ancient temples to more modern examples like Hotel Tassel in Brussels (Victor Horta, architect, 1893) and the structures of Gare do Oriente in Lisbon (Santiago Calatrava, architect, 1998). More intriguing still is the architectural expression of mathematical proportions or arrangements that occur in nature, the meaning of which has been fodder for philosophical prose since Aristotle and Euclid. Many cultures have used these mathematical relationships in the construction of buildings and sacred spaces. The Egyptian Pyramids, the Parthenon (447–438 BC), Notre Dame in Paris (beginning in 1163), the Taj Mahal in India (1632–1653), the CN Tower in Toronto (1976), and the Eden Project Education Centre in Corwall, UK (2000), are all alleged to exhibit the Golden Mean. Design Objectives and Guidance. The objective of Biomorphic Forms and Patterns is to provide representational design elements within the built environment that allow users to make connections to nature. The intent is to use biomorphic forms and patterns in a way that creates a more visually preferred environment that enhances cognitive performance while helping reduce stress. Humans have been decorating living spaces with representations of nature since time immemorial, and architects have long created spaces using elements inspired by trees, bones, wings, and seashells. Many classic building ornaments are derived from natural forms, and countless fabric patterns are based on leaves, flowers, and animal skins. Contemporary architecture and design have introduced more organic building forms with softer edges or even biomimetic qualities. There are essentially two approaches to applying Biomorphic Forms and Patterns, as either a cosmetic decorative component of a larger design or as integral to the structural or functional design. Both approaches can be utilized in tandem to enhance the biophilic experience. Design guidance translated from scientific research to help create a quality biomorphic condition includes: • Consideration of biomorphic forms and assemblies early in the design process to help

Biophilic Design

Biophilic Design, Fig. 12 Connection with natural systems. Outside the eighth floor penthouse studio of COOKFOX Architects in New York City sits a 3000 ft2 extensive green roof that changes color and vibrancy from season to season. Witnessing a hawk preying upon a small bird shifted employee perception of the green roof from being a decorative garden to a living ecosystem. This perception was reinforced when employees noticed changes in bee colony behavior during times of extreme

65

heat and humidity, when the honeybee apiary was invaded by robber bees, and again when the summer honey harvest looked and tasted different than the autumn harvest. After 10 years, the company was so attached to the experience afforded by the elevated ecosystem that they hauled the green roof materials with them when they transferred to a different building in 2017 (Credit: William D. Browning for Terrapin Bright Green)

66

Biophilic Design, Fig. 13 Biomorphic forms and patterns. The Art Nouveau Hotel Tassel (Victor Horta, architect, 1893) in Brussels, Belgium, is a favorite example of Biomorphic Forms and Patterns. The interior space in particular is rife with natural analogues, with graphic vine-like tendrils painted on the wall and designed into the banisters and railings, floor mosaics, window details, furniture, and columns. The curvaceous tiered steps seem to make distant reference to shells or flower petals (Credit: Henry Townsend)

establish spatial properties and other characteristics of the space. • Application of biomorphic patterns on 2 or 3 planes or dimensions (e.g., floor plane and wall, furniture, windows and soffits) for greater diversity and frequency of exposure (Fig. 13). Pattern 9. Material Connection with Nature A Material Connection with Nature is ensured through materials and elements from nature that, through minimal processing, reflect the local ecology or geology to create a distinct sense of place. A biophilic experience of a Material Connection with Nature feels rich, warm, and authentic, and sometimes stimulating to the touch. Roots of the Pattern. While scientific documentation on the health impact of natural materials is limited, available research is beginning to

Biophilic Design

shed light on opportunities for informed design. As such, the Material Connection with Nature pattern has evolved from a limited body of scientific research on physiological responses to variable quantities of natural materials, and the impact of natural color palette, particularly the color green, has on cognitive performance. One such study demonstrated that a difference in wood ratio on the walls of an interior space led to different physiological responses [66]. The researchers observed that a room with a moderate ratio of wood (i.e., 45% coverage) offered a more subjective “comfortable” feeling and exhibited significant decreases in diastolic blood pressure and significant increases in pulse rate, whereas a decrease in brain activity was observed in spaces with large ratios of wood (i.e., 90% coverage), which could be either highly restorative in a spa or doctor’s office or counterproductive in a space where high cognitive functionality is expected. In a series of four experiments examining the effect of the presence of the color green on the psychological functioning of participants, the results concluded that exposure to the color green before conducting a task “facilitates creativity performance, but has no influence on analytical performance” [67]. Humans are also able to distinguish more variations in the color green than of any other color. However, which variation(s) of the color green most influence creativity or other mind-body responses is not well understood. Design Objectives and Guidance. The objective of the Material Connection with Nature pattern is to explore the characteristics and quantities of natural materials optimal for engendering positive cognitive or physiological responses. In some cases, there may be several layers of information in materials that enhance the connection, such as learned knowledge about the material, familiar textures, or nested fractals that occur within a stone or wood grain pattern. Natural materials can be decorative or functional and are typically processed or extensively altered (e.g., wood plank, granite countertop) from their original “natural” state, and while they may be extracted from nature, they are only analogous of the items in their “natural” state.

Biophilic Design

Design guidance translated from scientific research to help create a quality Material Connection with Nature includes: • Specification of natural materials and colors is in quantities based on intended function of the space (e.g., to restore versus to stimulate). Variability of materials and applications is likely to be more effective than high ratios of any one material or color. • Selection of natural materials is preferred over synthetic variations; our brain receptors can tell the difference between real and synthetic nature. • Incorporation of the color green may help enhance creative environments; however, scientific studies on the impact of the color green have mostly been conducted in controlled lab environments, so dependence on color to engender creativity should be considered experimental (Fig. 14). Pattern 10. Complexity and Order Complexity and Order is rich sensory information that adheres to a spatial hierarchy similar to those encountered in nature. A biophilic experience of Complexity and Order feels engaging and information-rich, as an intriguing balance between boring and overwhelming. Roots of the Pattern. The Complexity and Order pattern has evolved from research on fractal geometries and preferred views [68–71]; the Biophilic Design, Fig. 14 Material connection with nature. The suites at the One Hotel Brooklyn Bridge in Brooklyn, New York, are designed with wood, leather, and textured carpet to deliberately connect guests to the park outside the window (Credit: Allison Bernett for Terrapin Bright Green)

67

perceptual and physiological stress responses to the complexity of fractals in nature, art, and architecture [65, 68, 69]; and the predictability of the occurrence of design in nature [105]. Research has repeatedly confirmed correlations between fractal geometries in nature and those in art and architecture [65, 69], but there are opposing opinions about which fractal dimension is optimal for engendering a positive health response, whether an optimal ratio exists, or if such a ratio is even important to identify as a design metric or guideline. Nikos Salingaros [68] has examined a series of these perspectives with great clarity, noting that the range of preferred fractal dimensions is potentially quite broad (D = 1.3–1.8) depending on the application [68]. Nested fractal designs expressed as a third iteration of the base design (i.e., with scaling factor of 3, see illustration) are more likely to achieve a level of complexity that conveys a sense of order and intrigue and reduces stress [68]. This quality is lost in much of modern architecture, which tends to limit complexity to the second iteration, and consequently results in an arguably dull and inadequately nurturing form that fails to stimulate the mind or engender physiological stress reduction. At either end of the spectrum, both nonfractal artwork and high-dimensional fractal artwork have been shown to induce stress [69, 70]. Overly complex designs and environments may result in psychological stress and even nausea. According

68

to Judith Heerwagen and Roger Ulrich, occupants in a US Navy office in Mississippi reported nausea, headaches, and dizziness, symptoms frequently associated with poor indoor air quality or poor ventilation due to the interaction of patterns. Multiple wallpaper patterns, complex patterns in carpets, and moiré patterns in seating fabrics caused surfaces to appear to move as occupants walked through the space and therefore caused extreme visual perception problems. Fractal patterns can be identified in classical art and vernacular architecture from the column capitals of ancient Greece and Egypt, to the art of Ancient Mayans and the Eiffel Tower in Paris (1889). Fractals are also evident in such wellknown works as those of Botticelli, Vincent van Gogh, and Jackson Pollock. Design Objectives and Guidance. The objective of the Complexity and Order pattern is to provide symmetries and fractal geometries, configured with a coherent spatial hierarchy, to create a visually nourishing environment that engenders a positive psychological or cognitive response [68]. Fractals can exist at any scale, from desktop trinkets or textile patterns, to façade design, to a city grid or regional transport infrastructure. Scenes in nature typically support multiple fractal dimensions – savanna landscapes often support mid-range fractal dimensions – so there are potentially many opportunities to incorporate fractals. A familiar challenge in the built environment is in identifying the balance between an information-rich environment that is interesting and restorative, and one with an information surplus that is overwhelming and stressful. Targeting an optimal dimensional ratio for design applications can be problematic (i.e., time consuming, inconsistent, and even inaccurate), of questionable long-term value, and essentially less important than the incorporation of thirditeration fractal designs. As Salingaros [68] points out, identifying precise fractal geometries in existing nature scapes, structures, and artworks is a challenge, whereas generating new works with complex fractals is quite easy, so specifying fractal artwork, for instance, may not always be the most efficient use of project resources [68].

Biophilic Design

Design guidance translated from scientific research to help create a quality Complexity and Order condition includes: • Prioritization of artwork and material selection, architectural expressions, and landscape and master planning schemes can help reveal fractal geometries and hierarchies. • Incorporation of fractal structures with scaled iterations of three are more likely to be perceived as interesting to the viewer than a design limited to two iterations. Underutilization of fractals in design could result in complete predictability and disinterest. • Use of algorithms of mathematical and geometric functions can produce fractal designs for architectural and planning applications with ease. Geometries with a mid-range dimensional ratio (broadly speaking, D = 1.3–1.75) should be targeted when a new fractal design is being created. Over-use of and/or extended exposure to high-fractal dimensions (D = 1.8–2.0) could instill discomfort or even fear, countering the intended response: to nourish and reduce stress. • Consideration for a new building or landscape design should take into account its impact on the fractal quality of the existing urban skyline (Fig. 15).

Pattern 11. Prospect Prospect is an unimpeded view over a distance for surveillance and planning. A biophilic experience of a good Prospect condition feels open and freeing, yet imparts a sense of safety and control, particularly when alone or in unfamiliar environments. Roots of the Pattern. The Prospect pattern is derived from visual preference research and spatial habitat responses, as well as cultural anthropology, evolutionary psychology [61], and architectural analysis. Health benefits are suggested to include reduced stress; reduced boredom, irritation, fatigue [73], and perceived vulnerability [74, 75]; as well as improved comfort [72, 76]. In evolutionary psychology terms, humans prefer habitats that are similar to the African savannas on which humankind evolved as a species. This

Biophilic Design

Biophilic Design, Fig. 15 Complexity and order. Tucked in between buildings of downtown Toronto, Ontario, is the Allen Lambert Galleria and Atrium at Brookfield Place. The cathedral-like structure designed by Santiago Calatrava (1992) is information rich, yet protecting, with its orderly columns that rise up into a canopy of complex tree-like forms, showers diffuse light and shadow onto the courtyard, and keeps visitors awestruck and engaged (Credit: Maarten van den Heuvel (creative commons))

becomes clear in visual preference research starting with Jay Appleton’s Experience of Landscape in 1977 [13], where he asked why certain views from the same vantage point are preferred over others. Kellert and Wilson [5] argue that view preferences, and possibly aesthetic preferences, have roots in referential points that benefit survival [5]. For example, flowers are indicators of healthy plant growth and to signal the availability of resources in the future [11]. The savanna, with its open terrain and copses of shade trees, becomes more favorable when combined with water, an understory of flowers and forbs, calm grazing animals, and evidence of human habitation. Distant prospect (>100 ft, >30 m) is preferred over shorter focal lengths (14 ft), a more drastic canopy differential may be necessary to achieve the desired outcome; freestanding or built-in alcoves and mezzanines are often effective. • Provision of more than one kind of refuge space – to address varying needs of larger populations or multiple activity types – through differing spatial dimensions, lighting conditions, and degrees of concealment. • Consideration for how a unique lighting control strategy may broaden the multifunctionality of a refuge space (Fig. 17). Pattern 13. Mystery Mystery is the promise of more information achieved through partially obscured views or other sensory devices that entice the individual to travel deeper into the environment. A biophilic experience of a Mystery condition has a palpable sense of anticipation, or of being teased, offering the senses a kind of denial and reward that compels one to further investigate the space. Roots of the Pattern. The Mystery pattern is largely based on the idea posited by Kaplan and

72

Kaplan [17] that people have two basic needs in environments: to understand and to explore and that these should occur “from one’s current position” in order to engender a sense of mystery. The Mystery pattern has evolved from research on visual preference and perceived danger [76, 110, 111] and is supported by research on pleasure responses to anticipatory situations [78–80]. Mystery engenders a strong pleasure response within the brain that may be a similar mechanism to that of anticipation, which is hypothesized to be an explanation for why listening to music is so pleasurable – in that assumptions are made as to what may be around the corner [78, 80]. The benefits of mystery conditions are suggested to include improved preference for a space; heightened curiosity; increased interest in gaining more information and greater likelihood of encountering other biophilic conditions. A quality mystery condition does not engender a fear response; the conditions that differentiate between surprise (i.e., fear) and pleasure center around the visual depth of field. An obscured view with a shallow depth of field has shown to lead to unpleasant surprises, whereas greater visual access with a medium (20 ft) to high (100 ft) depth of field is preferred [76]. A good mystery condition could also be expressed through the obscuring of the boundaries and a portion of the focal subject (i.e., room, building, outdoor space, or other information source), thereby enticing the user to anticipate the full extent of the subject and explore the space further [79]. Design Objectives and Guidance. Mystery characterizes a place where an individual feels compelled to move forward to see what is around the corner; it is the partially revealed view ahead. The objective of the Mystery pattern is to provide a functional environment that encourages exploration in a manner that supports stress reduction and cognitive restoration. While other “Nature of the Space” patterns can be experienced in a stationary position, mystery implies movement and analysis starting from a place perceived in a fundamentally positive way. Mystery conditions have their place among indoor and outdoor plazas, corridors, pathways,

Biophilic Design

parks, and other transitory spaces. The sense of mystery can be diluted over time and with routine exposure; however, strategies that include revolving content or information, such as peek-a-boo windows into common areas where activity is constantly changing, will be most effective in spaces routinely occupied by the same group of people. Design guidance translated from scientific research to help create a quality Mystery condition includes: • Incorporation of curving edges (rather than sharp corners) that slowly reveal a people is drawn through a space. • Integration of dramatic shading and shadow casting, while minimizing intensely dark shadows or shallow depths of field that could engender undesired surprise or fear. • Consideration for how the speed at which people are moving through a space may influence both the necessary size of a view aperture and the size of the focal subject, with faster speeds typically requiring bigger apertures. • Accommodation for organically evolved mystery conditions (e.g., low maintenance parks with winding paths) that change over time. Changes (e.g., vegetation height or density) should be monitored as they may enhance the mystery condition or degrade it as it evolves into a surprise condition (e.g., overgrowth of plantings leads to obscuring of depth of field) (Fig. 18). Pattern 14. Risk/Peril Risk/Peril is an identifiable threat coupled with a reliable safeguard. A biophilic experience of a Risk/Peril condition feels exhilarating and with an implied threat, maybe even a little mischievous or perverse. One feels that it might be dangerous, but intriguing, worth exploring, and possibly even irresistible. Roots of the Pattern. Risk can be generated by a learned or bio-phobic response triggered by a near and present danger. This danger, however, is inert and unable to cause harm due to a trusted element of safety. The defining difference between Risk/Peril and fear is the level of perceived threat and perceived control [112].

Biophilic Design

Biophilic Design, Fig. 17 Refuge. Set back from the street, Paley Park provides respite from the bustling activity and heat of midtown Manhattan. Designed in 1969 by Zion and Breen, the 2400 square foot pocket park offers visitors obscured visual access from the street, movable

73

chairs to allow flexibility in how much of the city street can be seen, high backed benches that face inward, a tree canopy to protect overhead that changes in density over the seasons, and a waterfall that masks both private conversations and street noise (Credit: William D. Browning)

Biophilic Design, Fig. 18 Mystery. The playful and daylit spiral staircases at Antinori Winery in Tuscany, Italy, entice visitors to explore each level of the facility with rewarding views across the vineyard (Credit: Tom Godber)

74

Having an awareness of a controllable risk can support positive experiences [113] that result in strong dopamine or pleasure responses [81, 82]. These experiences play a role in developing risk assessment during childhood [114, 115]. In adults, short doses of dopamine support motivation, memory, problem solving, and fight-or-flight responses, whereas long-term exposure to intense Risk/Peril conditions may lead to over-production of dopamine, which is implicated in depression and mood disorders. Design Objectives and Guidance. The objective of the Risk/Peril pattern is to arouse attention and curiosity, and refresh memory and problem solving skills. There are different degrees of risk that can be incorporated into a design depending on the intended user or the space available; a cantilevered walkway over a sheer cliff is an extreme case; viewing a predator in a zoo exhibit may provide risk with a greater sense of control, whereas rock-hopping through a gentle water feature presents the risk of getting wet. Design guidance translated from research to help create a quality Risk/Peril condition includes: • Incorporation of perceptions of risk in the project as early as concept or schematic design phases of the design process. Architectural solutions for unique spatial conditions are more difficult and costly to incorporate later in the design process. • Protection provided by the “element of safety” adequately protects the user from harm while still permitting a meaningful experience of risk. • Consideration for the target user audience to ensure a Risk/Peril design intervention is appropriate (Fig. 19).

Conceptual Issues for the Application of Biophilic Design Planning for Implementation

Increasingly dense urban environments, coupled with rising land values, elevate the importance of biophilic design across a spatial continuum from new and existing buildings, to parks and streetscapes and to campus, urban and regional planning. Each context supports a platform for myriad

Biophilic Design

Biophilic Design, Fig. 19 Risk/Peril. The spiral ramp at the Reichstadt in Berlin, Germany, lures visitors up and around with glass partitions on either side, providing extensive views in every direction. Some visitors get a thrill from walking higher and looking down, while others would rather stay at the base looking up at everyone else and, in doing so, get an equally exhilarating response to witnessing the risk taken by others (Credit: Catherine O. Ryan)

opportunities for integrating biophilic design and healthy building practices for people and society. Key perspectives have been identified to help focus the planning and design processes. Identifying desired responses and outcomes. It is vital for a designer to understand a project’s design intent, specifically, what the health or performance priorities are for the intended users. To identify design strategies and interventions that restore or enhance well-being, project teams should understand the health baseline or performance needs of the target population. One approach is to ask: what is the most biophilic space that can conceivably be designed? Another is to ask: how can biophilic design improve performance metrics already used by the client (e.g., company executives, school board, city officials),

Biophilic Design

such as absenteeism, perceived comfort, health care claims, asthma, ticket sales, or test scores. Given that many human responses to design are integral (e.g., reducing physiological stress and improving overall mood) and that there are countless combinations of design patterns and interventions, understanding health related priorities will help focus the design process. Health outcomes associated with biophilic spaces are of interest to building and portfolio managers and human resources administrators, because they inform long-term design and measurement best practices, and to planners, policy makers, and others because they inform public health policy and urban planning. Design strategies and interventions. Biophilic design patterns are flexible and replicable strategies for enhancing the user experience that can be implemented under a range of circumstances. Just as lighting design for a classroom will be different than for a spa or home library, biophilic design interventions are based on the needs of a specific population in a particular space and are likely to be developed from a series of evidence-based biophilic design patterns, ideally with a degree of monitoring and evaluation for efficacy. For example, a project team may embrace a “Visual Connection with Nature” to enhance the workplace experience for a series of interior fitouts for a portfolio of offices. The strategy would be to improve views and bring plants into the space; the interventions may include installing a green wall, orienting desks to maximize views to outdoors, and initiating an employee stipend for desk plants. The detail, location, and the extent to which each of these interventions is implemented will differ for each of the offices in the portfolio. A project team charged with reducing stress among emergency room nurses at the local hospital may intervene by replacing the abstract art with landscape paintings on the walls of the staff room and installing a small garden and seating area in the adjacent interior courtyard. While this project also uses the Visual Connection with Nature pattern, the selected interventions specifically target stress reduction for emergency room nurses based on a shared space they utilize routinely. Diversity of design strategies. Patterns in combination tend to increase the likelihood of

75

health benefits of a space. Incorporating a diverse range of design strategies can accommodate the needs of various user groups from differing cultures and demographics and create an environment that is psycho-physiologically and cognitively restorative. For instance, vegetated spaces can improve an individual’s self-esteem and mood, while the presence of water can have a relaxing effect. Adding multiple biophilic strategies for the sake of diversity may backfire unless they are integrative and support a unified design intent. Quality vs. quantity of intervention. When planning for implementation, common questions recur, such as how much is enough and what makes a good design great. A high quality intervention may be defined by the richness of content, user accessibility, and as mentioned above, diversity of strategies. A single high quality intervention can be more effective and have greater restorative potential than several low quality interventions. Climate, cost, and other variables may influence or limit feasibility of certain interventions, but should not be considered an obstacle to achieving a high quality application. For example, multiple instances of “Prospect” with a shallow to moderate depth of field and limited information in the view shed may not be as effective (at prompting the desired response) as a single powerful instance of Prospect with a moderate to high depth of field and an information-rich view shed. Duration of exposure and frequency of access. Identifying the most appropriate duration of exposure to a pattern, or combination of patterns, can be difficult. The ideal exposure time is likely dependent upon the user and desired effect, but as a general guideline, empirical evidence shows that positive emotions, mental restoration, and other benefits can occur in as little as 5–20 min of immersion in nature [27, 28, 30]. When a long duration of exposure is not possible or desired, positioning biophilic design interventions along paths that channel high levels of foot traffic will help improve frequency of access. Consider too that microrestorative experiences – brief sensory interactions with nature that promote a sense of well-being – often

76

designed in response to space-restriction, are more readily implementable, replicable, and often more accessible than larger interventions; frequent exposure to these small interventions may contribute to a compounded restoration response. Questions abound on matters of duration of exposure and frequency of access, including how persistent mental restoration is over different terms of exposure to nature; whether improvements plateau or continue incrementally with more exposure; and whether combinations of design patterns can further optimize a biophilic experience. These questions and others are being explored as research continues on the intersections of neuroscience and design [116]. Locally Appropriate Design

No two places are the same; this presents both challenges and opportunities for creativity in the application of biophilic design patterns. Several key considerations are being utilized to help frame, prioritize, or influence decision making in the design process. Climate, ecology and the vernacular. Historically, humans have built shelters from locally available materials that reflected the regional ecology; form and function were in response to the topography and climate. Known as vernacular architecture, these buildings and constructed landscapes connect to the settings they inhabit. Use of local timber, climate responsive design and xeriscaping – using native, drought tolerant plants to create landscape designs that resemble the climate of the surrounding landscape – can each be effective strategies in designing for a resilient, biophilic experience. Whether rural or urban, not all natural or tempered environments are “green” in color, nor should they be. Desert species and terrain can be equally important in reinforcing a biophilic connection to place. Some habitats may engender a stronger positive response than others, but a small bio-diverse savanna-like scene will most likely be preferred over an expansive yet trackless sand desert, open ocean, or dark forest. Character and density: Rural, suburban and urban environments. In rural environments, human-nature interactions are abundant, and this

Biophilic Design

regular exposure to nature has restorative qualities that are sometimes taken for granted. Suburban settings are typically rife with intuitively applied biophilic design; the suburban yard with shade trees, grass, low shrubs, and beds of flowers is essentially an analogue of the African savanna. Porches and balconies offer more than just quaintness and real estate value; many suburban homes and urban rowhouses are raised 18 inches or more, creating a Prospect-Refuge condition with views from windows, stoops and porches. The potential human health benefits are undervalued in highdensity settings where residential towers with balconies are both limited and only available to high-paying tenants. Land in urban environments is limited and at a premium, so it may be unrealistic to replicate features more suitable to a rural environment in terms of scale or abundance. As such, biophilic design strategies will differ depending on the local geography, land availability and ownership, zoning and political will. For instance, San Francisco, with its high-density urban form, implemented a “parklet” system, whereby temporary pop-up parks occupy parking spaces for limited periods of time [117]. In the narrow streets of Vienna, Austria, restaurants rent parking spaces for the entire summer and set up tables and temporary landscaping to provide outdoor dining. This brings nature into the urban core and within walking distance to a greater number of people, opening up the possibility for micro-restorative experiences and a reclamation of underutilized space for people. A different approach to integrating natural systems with urban systems is exhibited in Singapore’s “Skyrise Greenery” program. Given the high levels of development in tropical Singapore over the last 25 years – a period which saw the country’s populations grow by 2 million people – the government offered an incentive program to offset the loss of habitat, to increase interaction with natural stimuli, and to create the “City within a Garden.” This incentive program offers up to 75% of the costs for installing living roofs and walls (exterior and interior) for new construction [118]. Critically, the strategy must be integrative and appropriate to the character

Biophilic Design

and density of the place, and not just another word for ecosystem restoration that does not reflect the human biological relationship with nature. Scale and feasibility. Biophilic design patterns should be scaled to the surrounding environment and to the predicted user population for the space. Patterns can be applied at the scale of a microspace, a room, a building, a neighborhood or campus, and even an entire district or city. Each of these spaces will present different design challenges depending on the programming, user types and dynamics, climate, culture, and various physical parameters, as well as existing or needed infrastructure. Size and availability of space are two of the most common factors influencing feasibility of biophilic design patterns. For instance, the Prospect pattern (Pattern 11) typically requires significant space. Other patterns, such as Connection with Natural Systems (Pattern 7), may be most feasible where there is direct access to outdoor space, a challenge in dense urban environments. Yet small scale, microrestorative Visual and Nonvisual Connections with Nature (Patterns 1 and 2) and Presence of Water (Pattern 5) can also be very effective. For instance, the psychological benefits of nature actually have been shown to increase with exposure to higher levels of biodiversity [96], yet these benefits do not necessarily increase with greater natural vegetative area. A general understanding has since emerged that small yet bio-diverse microrestorative design interventions can be effective at engendering a restorative biophilic experience. Microrestorative might include moments of sensory contact with nature through a window, television, image, painting, or an aquarium. In urban environments where sensory overload is common [65], such experiences will be most valued and impactful when situated in locations with high foot traffic, allowing for a greater frequency of access to trigger the desired biophilic response. Traditional Japanese doorway gardens are a perfect example of replicable small-scale interventions. The speed at which one moves through an environment, whether rural or urban, impacts the level of observable detail and the perceived scale

77

of buildings and spaces. The General Motors “Tech Center” in Warren, Michigan, designed by architect Eero Saarinen in 1949, is designed to be experienced at 30 mph, so for the pedestrian, the scale seems oversized and the spacing of buildings is oddly far apart. This is why stores along strip malls have large, simple façades and signage, whereas stores within pedestrian zones tend to have smaller and perhaps more intricate signage. Similarly, the landscaping along freeway and highway greenbelts is typically done in large swaths for instant interpretability. In contrast, a pedestrian focused environment will have more fine-grained details in the landscape design to allow for pause, exploration, and a more intimate experience. Some patterns, such as Mystery and Risk/Peril (Patterns 13 and 14), might not be as feasible or cost-effective in an interior fit-out project because of the amount of space required to effectively implement the pattern. On the other hand, interior fit-outs are an excellent opportunity to introduce Natural Analogue patterns which can be applied to surfaces like walls, floors, and ceilings as well as furniture and window treatments. Not all aspects of biophilia are space dependent. Some patterns (e.g., Patterns 2, 4, 6, 7) are more visceral or temporal, requiring little to no floor area, and other patterns (e.g., Patterns 8–10) may simply guide design choices that were already a part of the design process. Major renovations, new construction, and master planning provide more opportunities for incorporating biophilic design patterns that are coupled with systems integration at the building, campus, or community scale. Culture and demographics. Current evolutionary hypotheses and theories state that contemporary landscape preferences are influenced by human evolution, reflecting the innate landscape qualities that enhanced survival for humanity through time. These schools of thought include the biophilia hypothesis [5, 10], the savanna hypothesis [11], the habitat theory and prospectrefuge theory [13], and the preference matrix [17]. While empirical research has shown that there is a degree of universality to landscape preferences among humans, preferences have been

78

modified by cultural influences, experiences, and socio-economic factors [119]. Variations in landscape preferences have thus emerged among ethnic groups, subcultures, genders, and age groups. Cultural constructs, social inertia, and ecological literacy suffuse differing perspectives on what constitutes natural, nature, wild, or beautiful [119, 120]. Environmental Generational Amnesia and the Ecological Aesthetic Theory help explain how some perspectives may have evolved, and these differences come to bear across countries and regions, as well as among neighborhoods within the same city. Cultures and groups across the world utilize landscapes and space in different ways [14]. Frequency of use, nature of use, participation rates, and purpose of visit all vary drastically between nationalities, cultures, and subgroups. These factors do not mean that certain ethnic groups have a lower appreciation for landscape or a less significant connection with nature. These groups simply utilize and interact with nature in ways that are compatible with their culture and needs. Identifying early on what those needs may be will help define parameters for appropriate design strategies and interventions. Age and gender are also known to influence biophilic response trends. Women report higher perceived levels of stress than men, yet are less likely than their male counterparts to use available natural outdoor vegetative space during the work day [121]. Of particular interest is that the degree of enhanced immune function due to immersion in nature has been observed to differ between the genders. For instance, following a forest walk, immune function was increased for a period of 30 days in men, but only 7 days in women [122], suggesting that interventions targeting female populations in the workplace may need to either prioritize indoor nature experiences or improve accessibility for prolonged outdoor nature experiences. Youth benefit the most from nature contact in terms of increasing self-esteem. The gains for selfesteem from nature contact are suggested to decline with age. Elderly and youth benefit the least in terms of mood enhancement from nature contact [30], yet both groups are equal in perceived restorative-ness of natural over urban environments [123]. With age also comes a differing

Biophilic Design

preference in landscape in regards to perceived safety. While an urban woodland may be an enticing place for adventure for a child or teenager, the same condition could be perceived as risky by adults and elderly populations [124], which could possibly be overcome by incorporating a Prospect-Refuge condition. Design Integration

Interdisciplinary planning and design. Developing an interdisciplinary strategy early on in a project will help ensure cost-effective opportunities are not lost before they are fully considered. Biophilia is but one piece of the puzzle to creating a vibrant, sustainable, and restorative environment. The integration of a multidisciplinary strategy in the early stages of development – through a stakeholder charrette process or similar – will put team members on equal footing and allow for the identification of potential strengths, challenges, and opportunities. In the long run, this approach will improve project satisfaction and save money. Biophilia as an environmental quality. Environmental quality is an umbrella term that refers to the sum of the properties and characteristics of a specific environment and how it affects human beings and other organisms within its zone of influence. Biophilia, like air quality, thermal comfort, and acoustics, is an essential component of environmental quality that expands the conversation from daylight, materials toxicity, and air, water and soil quality, to include human biological health and well-being. When integral to the environmental quality discussion, biophilia may also help dissolve the perceived division between human needs and building performance. Back-of-house and night shift workers are often the most deprived of biophilic experiences, while they are also the very people responsible for monitoring and maintaining building performance standards. From an architectural perspective, biophilic design patterns have the potential to refocus the designer’s attention on the links between people, health, high-performance design, and aesthetics. Multi-platform solutions. Thoughtful applications of biophilic design can create a multiplatform strategy for familiar challenges

Biophilic Design

traditionally associated with building performance such as thermal comfort, acoustics, energy, and water management, as well as larger scale issues such as asthma, biodiversity, and flood mitigation. Increased natural air flow can help prevent sick building syndrome; daylighting can cut energy costs in terms of heating and cooling [125]; and increased vegetation can reduce particulate matter in the air, reduce urban heat island effect, improve air infiltration rates, and reduce perceived levels of noise pollution [14]. These strategies can all be implemented in a manner that achieves a biophilic response for improved performance, health, and well-being. Biophilic design interventions that integrate with other building performance strategies have the potential to improve user experience and overall systems efficiency. Herbert Dreiseitl’s design for Prisma in Nürnberg, Germany, is a good example; sculptural water walls serve as both a thermal control device and exposed rainwater conduit, while contributing to the visual and acoustic ambiance of the enclosed garden-like atrium. For the design of the Khoo Teck Puat Hospital in Singapore, architect RMJM met with ecologists and engineers early in the project development process to employ biophilia, ecological conservation, and water-sensitive urban design to manage rainwater, mitigate loss of biodiversity, and create a restorative environment for patients, reaping more benefits for the project than any one of the three disciplines could have on their own [126]. The biophilic experiences are more likely to persist long term when they are embedded in the programming and infrastructure of a place. Controlling for effectiveness. Given that landscapes and people’s needs are in a constant state of flux, it is challenging to ensure the desired health response is always experienced. It is impossible to predict all future human-nature interactions or to ensure that the desired response recurs over a period of time for every user based on a particular strategy or intervention. There is an assumption that the efficacy of a biophilic experience is likely to rise and decline with diurnal and seasonal cycles. The health benefits of a view to nature may be diminished during winter months or completely negated for night shift workers when the view is shrouded in

79

darkness. Seasonal and indoor strategies are considered potential supporting solutions for maintaining a balance of biophilic experiences to help engender desired health responses throughout the year. User controls for lighting, heating, cooling, ventilation, and even noise can either complement design efforts or negate them when controls are mismanaged or underutilized – keeping the window blinds closed eliminates a Visual Connection with Nature, and adding high partitions in an open plan office eliminates opportunities for Prospect and a number of other patterns [127]. Behavior change is not often in the purview of the architect, so designing for controllability followed by occupant education may be critical. Maintenance of implemented strategies is also a consideration – will there be someone responsible for cleaning the fish tank and watering the plants? Providing training and resources for facility operators alongside a reference guide indicating appropriate maintenance requirements and parameters will help uphold the intended biophilic experience set forth in the design strategy. Tracking and measuring efficacy. Monitoring efficacy of implemented biophilic design patterns for the express purpose of improving health and well-being is a new branch of inquiry. Variability in the built environment, as discussed here, creates a challenging framework for verification; quantitative metrics are often desired but not always appropriate, and the highly invasive nature of some measurement techniques and tools (i.e., fMRI, EEG) adds a layer of complexity and cost. Many of the current techniques require strict control of variables and costs, which tends to limit the size of the test group. There are, however, several new technologies, like wristband monitors, and very light weight headband EEG that may open up new rapid methods of testing. Culture, climate, age, gender, landscape character, immigrant status, mental health, and genetic predispositions create a challenging labyrinth of data for comparison. Nevertheless, tracking and monitoring of human biological responses and outcomes triggered by a biophilic pattern is vital in the progress and further development of biophilic design as a best practice.

80

Standardization

Biophilic design is now a distinct and recognized area of design research and strategy, with growing commitment to documenting best practices and standardizing protocols for measurement and verification. Primary obstacles to standardization are health data privacy laws, variability in health baselines across populations, and consensus on quality across building typologies, climates, and populations. A recurring debate in the standardization process is how uniformly aspects of biophilic design can be meaningfully prioritized or required. The WELL Building Standard ® and the Living Building Challenge ® are two of the earliest examples of building assessment programs with qualitative and quantitative parameters for biophilic design implementation. Qualitative methods. Qualitative methods have emerged as a viable approach to assessing biophilic design based on research evidence around the presence of certain nature-based elements that can support specific human experiences: • Temporal and emotional experience (physical and spatial measurements; anecdotal feedback) • Perceived health impact (subjective data) • Physical presence and perceived intensity of biophilic characteristics Quantitative methods. Quantitative methods remain elusive due to skepticism and lack of consensus over the efficacy of imposing prescriptive measures for biophilic design. There remains an awareness that while biophilic design is applicable to almost any structure or place, the solutions for each space may vary and thus in practice must retain a degree of flexibility: • What constitutes “nature” tends to vary between rural, suburban, and urban populations, and across cultures. • Baseline health metrics vary across each target user groups. • Applied strategies vary depending on designer interpretation and user needs, both of which may mature or change over time. • Adaptability and relevance of quantitative methods vary across climates and ecosystems,

Biophilic Design

cultures, user demographics, and building typologies. • Streamlined processes for noninvasive measurement and verification of quantitative metrics after construction are not standardized or cost-effective.

Future Directions Researchers continue to refine the scientific basis for these biophilic design patterns and hypothesize about new ones such as Color or Awe. Color resurfaces periodically as having potential to be considered its own pattern, but climatic differences and cultural perceptions make it a particularly complex topic on which to provide science-based design guidance specific to health and wellbeing. Awe was once assumed to be experienced with the alignment of a handful of natural analogues and spatial patterns. However, research has begun to reveal that there may be something more profound at play. Awe may in fact be a spatial experience that engages multiple centers in the brain simultaneously to cause a sort of stimulation overload resulting in the rare but well-known physical reactions of “eye-popping” and “jaw-dropping.” Many proponents of biophilic design argue that the research is simply corroborating the rediscovery of the intuitively obvious. In the 12,000 years since humans began farming and other activities that transformed the natural landscape, only in the last 250 years have modern cities become our habitat. Today, with more people living in cities than in the countryside, the built environment is an inescapable reality of everyday life. Unfortunately, much modern design is devoid of any profound connection with nature. In coming decades, 70% of the world’s population is projected to be living in cities. With this impending shift, the need for designs to (re)connect people to an experience of nature becomes ever more important. The prevailing attitude is that while more robust empirical evidence may take time to accumulate, the industry, government, and the design professions should actively be restoring the human-nature connection in the built environment through research, education, and best practices.

Biophilic Design

Bibliography Primary Literature 1. Browning WD, Kallianpurkar NK, Ryan CO, Labruto L (2012) The economics of biophilia: why designing with nature in mind makes financial sense, 2015 edn. Terrapin Bright Green LLC, New York, p 40 2. US Department of Labor (2010) Absences from work of employed full-time wage and salary, US Department of Labor, Washington, DC 3. Bureau of Labor Statistics (2017) Employer costs for employee compensation new release, US Department of Labor, Washington, DC, June 2011 4. BOMA (1991) Benchmarking reports. Experience exchange report, BOMA International 5. Kellert SR, Wilson EO (1993) The biophilia hypothesis. Island Press, Washington, DC 6. Olmsted FL (1865) Introduction to Yosemite and the Mariposa Grove: A preliminary report 7. Kellert SR (2011) Director, biophilic design: the architecture of life [Film]. Bullfrog Films 8. Sternberg EM (2009) Healing spaces. Bleknap Harvard University Press, Cambridge, MA 9. Fromm E (1964) The heart of man. Harper & Row, New York 10. Wilson EO (1984) Biophilia. Harvard University Press, Cambridge, MA, p 157 11. Orians GH, Heerwagen JH (1992) Evolved responses to landscapes. In: Barkow JH, Cosmides L, Tooby J (eds) The adapted mind: evolutionary psychology and the generation of culture. Oxford University Press, New York, pp 555–579 12. Heerwagen JH, Orians GH Adaptations to windowlessness: a study of the use of visual decor in windowed and windowless offices. Environ Behav 18(5):623–639 13. Appleton J (1977, 1996) The experience of landscape, Rev. edn. Wiley, London/New York, pp xiv, 282 14. Forsyth A, Musacchio LR (2005) Designing small parks: a manual for addressing social and ecological concerns. Wiley, Hoboken, pp 13–30, 60–65, 74–82, 95–98 15. Balling JD, Falk JH (1982) Development of visual preference for natural environments. Environ Behav 14(1):5–28 16. Nassauer JI (1995) Messy ecosystems, orderly frames. Landsc J 14(2):161–169 17. Kaplan R, Kaplan S (1989) The experience of nature: a psychological perspective. Cambridge University Press, Cambridge 18. Kahn PH Jr, Severson RL, Ruckert JH (2009) The human relation with nature and technological nature. Curr Dir Psychol Sci 18(1):37–42 19. Browning WD, Romm JJ (1994) Greening the building and the bottom line. Rocky Mountain Institute, Snowmass 20. Ulrich RS (1984) View through a window may influence recovery from surgery. Science 224:420

81 21. Heerwagen JH, Hase B (2001) Building biophilia: connecting people to nature in building design. Environmental Design & Construction, March–April 22. Louv R (2008) Last child in the woods: saving our children from nature-deficit disorder. Algonquin Books, New York, p 390 23. Kellert SR, Heerwagen JH, Mador ML (eds) (2008) Biophilic design. Wiley, Hoboken 24. Browning WD, Ryan CO, Clancy JO (2014) 14 patterns of biophilic design: improving health and wellbeing in the built environment, 2015th edn. Terrapin Bright Green LLC, New York, p 60 25. Kellert SR, Calabrese EF (2015) The practice of biophilic design [Online]. Available: http://www. biophilic-design.com/. Accessed 2 June 2017 26. van den Berg AE, Hartig T, Staats H (2007) Preference for nature in urbanized societies: stress, restoration, and the pursuit of sustainability. J Soc Issues 63(1):79–96 27. Brown DK, Barton JL, Gladwell VF (2013) Viewing nature scenes positively affects recovery of autonomic function following acute-mental stress. Environ Sci Technol 47:5562–5569 28. Tsunetsugu Y, Miyazaki Y (2005) Measurement of absolute hemoglobin concentrations of prefrontal region by near-infrared time-resolved spectroscopy: examples of experiments and prospects. J Physiol Anthropol Appl Hum Sci 24(4):469–472 29. Biederman I, Vessel E (2006) Perceptual pleasure & the brain. Am Sci 94(1):249–255 30. Barton J, Pretty J (2010) What is the best dose of nature and green exercise for improving mental health. Environ Sci Technol 44:3947–3955 31. Park BJ, Tsunetsugu Y, Kasetani T, Morikawa T, Kagawa T, Miyazaki Y (2009) Physiological effects of forest recreation in a young conifer forest in Hinokage Town, Japan. Silva Fennica 43(2):291–301 32. Hartig T, Evans GW, Jamner LD, Davis DS, Gärling T (2003) Tracking restoration in natural and urban field settings. J Environ Psychol 23:109–123 33. Orsega-Smith E, Mowen AJ, Payne LL, Godbey G (2004) Interaction of stress and park use on psycho-physiological health in older adults. J Leis Res 36(2):232–256 34. Ulrich RS, Simons RF, Losito BD, Fiorito E, Miles MA, Zelson M (1991) Stress recovery during exposure to natural and urban environments. J Environ Psychol 11:201–230 35. Mehta R, Zhu R, Cheema A (2012) Is noise always bad? Exploring the effects of ambient noise on creative cognition. J Consum Res 36(4):784–799 36. Ljungberg J, Neely G, Lundström R (2004) Cognitive performance and subjective experience during combined exposures to whole-body vibration and noise. Int Arch Occup Environ Health 77:217–221 37. Li Q, Kobayashi M, Inagaki H, Wakayama Y, Katsumata M, Hirata Y, Li Y, Hirata K, Shimizu T, Nakadai A, Kawada T (2012) Effect of phytoncides from forest environments on immune function. Forest Med:157–167

82 38. Jahncke H, Hygge S, Halin N, Green AM, Dimberg K (2011) Open-plan office noise: cognitive performance and restoration. J Environ Psychol 31:373–382 39. Tsunetsugu Y, Park BJ, Miyazaki Y (2010) Trends in research related to “Shinrin-yoku” (taking in the forest atmosphere or forest bathing) in Japan. Environ Health Prev Med 15:27–37 40. Kim JT, Ren CJ, Fielding GA, Pitti A, Kasumi T, Wajda M, Lebovits A, Bekker A (2007) Treatment with lavender aromatherapy in the post-anesthesia care unit reduces opioid requirements of morbidly obese patients undergoing laparoscopic adjustable gastric banding. Obes Surg 17(7):920–925 41. Stigsdotter UA, Grahn P (2003) Experiencing a garden: a healing garden for people suffering from burnout diseases. J Ther Hortic 14:38–48 42. Kahn PH, Jr BF, Gill B, Hagman J, Severson RL, Freier NG, Feldman EN, Carrere S, Stolyar A (2008) A plasma display window? The shifting baseline problem in a technology mediated natural world. J Environ Psychol 28(1):192–199 43. Park BJ, Tsunetsugu Y, Ishii H (2008) Physiological effects of Shinrin-yoku (taking in the atmosphere of the forest) in a mixed forest in Shinano Town, Japan. Scand J For Res 23:278–283 44. Beauchamp MS, Lee KE, Haxby JV, Martin A (2003) FMRI responses to video and point-light displays of moving humans and manipulable objects. J Cogn Neurosci 15(7):991–1001 45. Li Q, Kobayashi M, Wakayama Y, Inagaki H, Katsumata M, Hirata Y, Hirata K, Shimizu T, Kawada T, Park J, Ohira T, Kagawa T, Miyazak Y (2009) Effect of phytoncide from trees on human natural killer cell function. Int J Immunopathol Pharmacol 22(4):951–959 46. Windhager S, Atzwangera K, Booksteina FL, Schaefera K (2011) Fish in a mall aquarium-an ethological investigation of biophilia. Landsc Urban Plan 99:23–30 47. Tham KW, Willem HC (2005) Temperature and ventilation effects on performance and neurobehavioral – related symptoms of tropically acclimatized call center operators near thermal neutrality. ASHRAE Trans:687–698 48. Wigö H (2005) Technique and human perception of intermittent air velocity variation, Centre for Built Environment 49. Hartig T, Mang M, Evans GW (1991) Restorative effects of natural environment experience. Environ Behav 23:3–26 50. Zhang H, Arens E, Huizenga C, Han T (2010) Thermal sensation and comfort models for nonuniform and transient environments: part II: local comfort of individual body parts. Build Environ 45(2):389–398 51. Arens E, Zhang H, Huizenga C (2006) Partial- and whole-body thermal sensation and comfort, part II: non-uniform environmental conditions. J Therm Biol 31:60–66

Biophilic Design 52. de Dear R, Brager G (2002) Thermal comfort in naturally ventilated buildings. Energ Buildings 34:549–561 53. Parkinson T, de Dear R, Candido C (2012) Perception of transient thermal environments: pleasure and alliesthesia. In: Proceedings of 7th Windsor Conference, Windsor 54. Alvarsson J, Wiens S, Nilsson M (2010) Stress recovery during exposure to nature sound and environmental noise. Int J Environ Res Public Health 7(3):1036–1046 55. Pheasant RJ, Fisher MN, Watts GR, Whitaker DJ, Horoshenkov KV (2010) The importance of auditory – visual interaction in the construction of ‘Tranquil Space’. J Environ Psychol 30:501–509 56. Hunter MD, Eickhoff SB, Pheasant RJ, Douglas MJ, Watts GR, Farrow TFD, Hyland D, Kang J, Wilkinson ID, Horoshenkov KV, Woodruff PWR (2010) The state of tranquility: subjective perception is shaped by contextual modulation of auditory connectivity. NeuroImage 53:611–618 57. White M, Smith A, Humphryes K, Pahl S, Snelling D, Depledge M (2010) Blue space: the importance of water for preference, affect and restorativeness ratings of natural and built scenes. J Environ Psychol 30(4):482–493 58. Karmanov D, Hamel R (2008) Assessing the restorative potential of contemporary urban environment (s). Landsc Urban Plan 86:115–125 59. Ruso B, Atzwanger K (2003) Measuring immediate behavioural responses to the environment. Michigan Psychol 4:12 60. Ulrich RS (1983) Aesthetic and affective response to natural environment. In: Altman I, Wohlwill JF (eds) Behavior and the natural environment. Plenum Press, New York, pp 85–125 61. Heerwagen JH, Orians GH (1993) Humans, habitats and aesthetics. In: Kellert SR, Wilson EO (eds) The biophilia hypothesis. Island Press, Washington, DC, pp 138–172 62. Figueiro MG, Brons JA, Plitnick B, Donlan B, Leslie RP, Rea MS (2011) Measuring circadian light and its impact on adolescents. Light Res Technol 43(2):201–215 63. Beckett M, Roden LC (2009) Mechanisms by which circadian rhythm disruption may lead to cancer. S Afr J Sci 105:415–420 64. Kim SY, Kim JJ (2007) Effect of fluctuating illuminance on visual sensation in a small office. Indoor Built Environ 16(4):331–343 65. Joye Y (2007) Architectural lessons from environmental psychology: the case of biophilic architecture. Rev Gen Psychol 11(4):305–328 66. Tsunetsugu Y, Miyazaki Y, Sato H (2007) Physiological effects in humans induced by the visual stimulation of room interiors with different wood quantities. J Wood Sci 53(1):11–16 67. Lichtenfeld S, Elliot AJ, Maier MA, Pekrun R (2012) Fertile green: green facilitates creative performance. Personal Soc Psychol Bull 38(6):784–797

Biophilic Design 68. Salingaros NA (2012) Fractal art and architecture reduce physiological stress. J Biourbanism 2(2):11–28 69. Taylor RP (2006) Reduction of physiological stress using fractal art and architecture. Leonardo 39(3):245–251 70. Hägerhäll CM, Laike T, Taylor RP, Küller M, Küller R, Martin TP (2008) Investigations of human EEG response to viewing fractal patterns. Perception 37:1488–1494 71. Hägerhäll CM, Purcella T, Taylor R (2004) Fractal dimension of landscape silhouette outlines as a predictor of landscape preference. J Environ Psychol 24:247–255 72. Grahn P, Stigsdotter UK (2010) The relation between perceived sensory dimensions of urban green space and stress restoration. Landsc Urban Plan 94:264–275 73. Clearwater YA, Coss RG (1991) Functional esthetics to enhance wellbeing. In: Harrison C, McKay C (eds) From Antarctica to outer space. Springer, New York, p 410 74. Petherick N (2000) Environmental design and fear: the prospect-refuge model and the University College of the Cariboo Campus. West Geogr 10(1):89–112 75. Wang K, Taylor RB (2006) Simulated walks through dangerous alleys: impacts of features and progress on fear. J Environ Psychol 26:269–283 76. Herzog TR, Bryce AG (2007) Mystery and preference in within-forest settings. Environ Behav 39(6):779–796 77. Ulrich RS (1993) Biophilia, biophobia and natural landscapes. In: Kellert SR, Wilson EO (eds) The biophilia hypothesis. Island Press, Washington, DC, pp 73–137 78. Salimpoor VN, Benovoy M, Larcher K, Dagher A, Zatorre RJ (2011) Anatomically distinct dopamine release during anticipation and experience of peak emotion to music. Nat Neurosci 14(2):257–264 79. Ikemi M (2005) The effects of mystery on preference for residential façades. J Environ Psychol 25:167–173 80. Blood A, Zatorre RJ (2001) Intensely pleasurable responses to music correlate with activity in brain regions. In: Proceedings from the National Academy of Sciences 81. Wang DV, Tsien JZ (2011) Convergent processing of both positive and negative motivational signals by the VTA dopamine neuronal populations. PLoS One 6(2):1–14 82. Zald DH, Cowan RL, Riccardi P, Baldwin RM, Ansari MS, Li R, Shelby ES, Smith CE, McHugo M, Kessler RM (2008) Midbrain dopamine receptor availability is inversely associated with novelty-seeking traits in humans. J Neurosci 28(53):14372–14378 83. Alcock I, White MP, Wheeler BW, Fleming LE, Depledge MH (2014) Longitudinal effects on mental

83

84.

85.

86.

87.

88. 89.

90. 91.

92.

93.

94. 95.

96.

97.

98.

99.

100.

health of moving to greener and less green urban areas. Environ Sci Technol 48(2):1247–1255 Kandel ER, Schwartz JH, Jessell TM, Siegelbaum SA, Hudspeth AJ (2013) Principles of neural science, 5th edn. McGraw Hill, New York Steg L (2007) Environmental psychology: history, scope & methods. In: Steg L, van den Berg AE, de Groot JIM (eds) Environmental psychology: an introduction, 1st edn. Wiley-Blackwell, Chichester, pp 1–11 Alexander C, Ishikawa S, Silverstein M, Jacobson M, Fiksdahl-King I, Angel S (1977) A pattern language: towns, buildings, construction. Oxford University Press, New York, p 1171 Kaplan R, Kaplan S, Ryan RL, (1998) With people in mind: design and management of everyday nature. Island Press, Washington, DC, pp 1–6, 67–107 Jacobson M, Silverstein M, Winslow B (2002) Patterns of home. The Taunton Press, Newtown Salingaros NA (2013) Unified architectural theory: form, language, complexity. Sustasis Foundation, Portland Salingaros NA (2000) The structure of pattern languages. Archit Res Q 4:149–162 Leather P, Pyrgas M, Beale D, Lawrence C (1998) Windows in the workplace: sunlight, view, and occupational stress. Environ Behav 30(6):739+ Bloomer K (2008) The problem of viewing nature through glass. In: Kellert SR, Heerwagen JH, Mador ML (eds) Biophilic design. Wiley, Hoboken, pp 253–262 van den Berg AE, Koole SL, van der Wulp NY (2003) Environmental preference and restoration: (how) are they related? J Environ Psychol 23:135–146 Hartig T (1993) Nature experience in transactional perspective. Landsc Urban Plan 25:17–36 Knopf RC (1987) Human behavior, cognition, and affect in the natural environment. In: Stokols D, Altman I (eds) Handbook of environmental psychology. Wiley, New York, pp 783–825 Fuller RA, Irvine KN, Devine-Wright P, Warren PH, Gaston KJ (2007) Psychological benefits of greenspace increase with biodiversity. Biol Lett 3(4):390–394 Lee KE, Williams KH, Sargent LD, Williams NSG, Johnson KA (2015) 40-second green roof views sustain attention: the role of micro-breaks in attention restoration. J Environ Psychol 42:182–189 Yamane K, Kawashima M, Fujishige N, Yoshida M (2004) Effects of interior horticultural activities with potted plants on human physiological and emotional status. Acta Hortic (639):37–43 Koga K, Iwasaki Y (2013) Psychological and physiological effect in humans of touching plant foliage – using the semantic differential method and cerebral activity as indicators. J Physiol Anthropol 32(1):7 de Dear R, Brager G, Cooper DD (1997) Developing an adaptive model of thermal comfort and preference.

84

101.

102.

103.

104.

105.

106.

107.

108.

109.

110.

111.

112.

113.

114. 115.

116.

117.

Biophilic Design Final Report on ASHRAE RP–884. MRL: Australia, pp 296 Mower GD (1976) Perceived intensity of peripheral thermal stimuli is independent of internal body temperature. J Comp Physiol Psychol 90(12):1152–1155 Kaplan S (1995) The restorative benefits of nature: toward an integrative framework. J Environ Psychol 15:169–182 Nicol JF, Humphreys MA (2002) Adaptive thermal comfort and sustainable thermal standards for buildings. Energ Buildings 34(1):563–572 Nicklas MH, Bailey GB (1996) Student performance in daylit schools [Online]. Available: http://www. innovativedesign.net/Profile-Resources-TechnicalPapers.html. Accessed June 2012 Bejan A, Zane JP (2012) Design in nature: how the constructal law governs evolution in biology, physics, technology, and social organization. Random House First Anchor Books, New York, p 304 Dosen AS, Ostwald MJ (2013) Prospect and refuge theory: constructing a critical definition for architecture and design. Int J Des Soc 6(1):9–24 Hildebrand G (1991) The Wright space: pattern & meaning in Frank Lloyd Wright’s houses. University of Washington, Seattle Nordh H, Hartig T, Hägerhäll CM, Fry G (2009) Components of small urban parks that predict the possibility for restoration. Urban For Urban Green 8:225–235 Ruddlell EJ, Hammitt WE (1897) Prospect refuge theory: a psychological orientation for edge effects in recreation environment. J Leis Res 19(4):249–260 Herzog TR, Kropscott LS (2004) Legibility, mystery, and visual access as predictors of preference and perceived danger in forest settings without pathways. Environ Behav 36:659–677 Nasar JL, Fisher B (1993) ‘Hot spots’ of fear and crime: a multi-method investigation. J Environ Psychol 13:187–206 Rapee R (1997) Perceived threat and perceived control as predictors of the degree of fear in physical and social situations. J Anxiety Disord 11:455–461 van den Berg AE, ter Heijne M (2005) Fear versus fascination: an exploration of emotional responses to natural threats. J Environ Psychol 25:261–272 Louv R (2009) Do our kids have nature-deficit disorder. Health Learn 67(4):24–30 P. H. Kahn Jr and S. R. Kellert, Children and nature: psychological, sociocultural, and evolutionary investigations, Cambridge, MA: MIT Press, 2002 Ryan CO, Browning WD, Clancy JO, Andrews SL, Kallianpurkar NB Biophilic design patterns: emerging nature-based parameters for health and wellbeing in the built environment. Int J Archit Res 8(2):62–76 C. o. S. Francisco (2013) San Francisco Parklet manual. San Francisco Planning Department, San Francisco, pp 1–12

118. Beatley T (2012) Singapore: city in a garden, 22 July [Online]. Available: http://biophiliccities.org/blogsingapore/ 119. Tveit MS, Sang AO, Hägerhall CM (2007) Scenic beauty: visual landscape assessment and human landscape perception. In: Steg L, van den Berg AE, de Groot JIM (eds) Environmental psychology: an introduction. Wiley, Chicester, pp 37–46 120. Zube EH, Pitt DG (1981) Cross-cultural perception of scenic and heritage landscapes. Landsc Plan 8:69–81 121. Lottrup L, Grahn P, Stigsdotter UK (2013) Workplace greenery & perceived level of stress: benefits of access to a green outdoor environment at the workplace. Landscape & Urban Planning 110(5):5–11 122. Li Q (2010) Effect of forest bathing trips on human immune function. Environ Health Prev Med 15(1):9–17 123. Berto R (2007) Assessing the restorative value of the environment: a study on the elderly in comparison with young adults and adolescents. Int J Psychol 42(5):331–341 124. Kopec D (2006) In: Kontzias OT (ed) Environmental psychology for design. Fairchild Publications Inc., New York, pp 38–57 125. Loftness V, Snyder M (2008) Where windows become doors. In: Kellert SR, Heerwagen JH, Mador ML (eds) Biophilic design. Wiley, Hoboken, pp 119–131 126. Health A (2013) Khoo Teck Puat Hospital [Online]. Available: http://www.ktph.com.sg/uploads/KTPH_ EBook/files/assets/basic-html/index.html#1. Accessed June 2014 127. Urban Green Council (2013) Seduced by the view: a closer look at all-glass buildings, Urban Green Council, USGBC-NY Chapter, New York 128. Elzeyadi IMK (2012) Quantifying the impacts of green schools on people and planet. Research presented at the USGBC Greenbuild Conference & Expo, San Francisco, pp 48–60 129. Heschong L (1979) Thermal delight in architecture. MIT Press, Cambridge, MA 130. Zhang H (2003) Human thermal sensation and comfort in transient and non-uniform thermal environments. PhD thesis, CEDR, University of California, Berkeley. http://escholarship.org/uc/item/11m0n1wt 131. Kaplan S (1988) Perception and landscape: conceptions and misconceptions. In: Nasar J (ed) Environmental aesthetics: theory, research, and applications. Cambridge University Press, Cambridge, UK, pp 45–55 132. Heerwagen JH (2006) Investing In People: The Social Benefits of Sustainable Design. Rethinking Sustainable Construction. Sarasota, FL, pp 19–22 133. Kohno M, Ghahremani DG, Morales AM, Robertson CL, Ishibashi K, Morgan AT, Mandelkern MA, London ED (2013) Risk-Taking Behavior: Dopamine D2/D3 Receptors, Feedback, and Frontolimbic Activity. Cerebral Cortex 25(1):236–245

Biophilic Design

Books and Reviews Bloomer K (2008) The problem of viewing nature through glass. In: Kellert SR, Heerwagen JH, Mador ML (eds) Biophilic design. Wiley, Hoboken, pp 253–262 Cooper R (1968) The psychology of boredom. Sci J 4(2):38–42 Heerwagen JH, Orians GH (1993) Humans, habitats and aesthetics. In: Kellert SR, Wilson EO (eds) The biophilia hypothesis. Island Press, Washington, DC, pp 138–172 Heijdens, Simon (2014) Shade. Installation. Web, 15 June 2017. https://vimeo.com/106669679 Heschong Mahone Group (1999) Daylighting in schools: an investigation into the relationship between daylighting and human performance. Pacific Gas and Electric Company: California Board for Energy Efficiency Third Party Program Heschong Mahone Group (2003) Windows and classrooms: a study of student performance and the indoor environment. Pacific Gas and Electric Company: California Board for Energy Efficiency Third Party Program Hordh H, Hartig T, Hägerhäll CM, Fry G (2009) Components of small urban parks that predict the possibility of restoration. Urban For Urban Green 8(4):225–235

85 Hosey L (2012) The shape of green: aesthetics, ecology, and design. Island Press, Washington, DC, p 216 Kuo FE, Taylor AF (2004) A potential natural treatment for attention-deficit/hyperactivity disorder: evidence from a national study. Am J Public Health 94(9):1580–1596 Leslie RP, Conway KM (2007) The lighting pattern book for homes. Rensselaer Polytechnic Institute, New York, p 222 Mehaffy MW, Salingaros NA (2015) Design for a living planet: settlement, science, and the human future. Sustasis Press, Portland Muir J (1877) Mormon Lilies. San Francisco Daily Evening Bulletin, 19 July 1877 Salingaros NA, Masden KG II (2008) Intelligence-based design: a sustainable foundation for worldwide architectural education. Int J Archit Res 2(1):129–188 Selhub EM, Logan AC (2012) Your brain on nature: the science of nature’s influence on your health, happiness, and vitality. Wiley, Mississauga Smithsonian Institute (2014) Human evolution timeline interactive. Web, 15 June 2017. http://humanorigins. si.edu/evidence/human-evolution-timeline-interactive Sosolimited (nd) Diffusion choir. Kinetic sculpture installation. Web, 15 June 2017. https://vimeo.com/ 187037469

Sustainable and Healthy Built Environment Vivian Loftness1 and Megan Snyder2 1 Center for Building Performance and Diagnostics, Carnegie Mellon University, Pittsburgh, PA, USA 2 School of Architecture and Center for Building Performance and Diagnostics, Carnegie Mellon University, Pittsburgh, PA, USA

Article Outline Glossary Definition of the Subject Introduction: Linking Definitions of Sustainability and Health CMU: Seven Principles for the Design of Sustainable Built Environments A Definition of Health to Be Integral with Sustainable Design Linking Health and the Built Environment Sustainable Land Use for Health Sustainable Building Massing and Enclosure for Health Sustainable HVAC for Health Sustainable Lighting for Health Sustainable Interior Systems for Health: Materials and Ergonomics Sustainable Maintenance and Operations for Health Calculating the Life Cycle Benefits of Sustainable Design and Health Health and the Built Environment: A Research Mandate Bibliography

Glossary BIDS™ The Building Investment Decision Support tool developed by the Center for Building Performance at Carnegie Mellon University to

both assemble research from around the world linking design decisions to performance outcomes and to create a “triple bottom line” calculator to support changes in design decision-making. Biophilia Introduced by E.O. Wilson, the biophilia hypothesis suggests that there is an instinctive bond between human beings and other living systems that must be met by building design that ensures critical connections. Cornell medical index An index and questionnaire created by Cornell in 1949 to consistently collect the breadth of pertinent medical and psychiatric data on patients given limited physician time. Epidemiological case studies Quantitative studies of a group of individuals in controlled environmental conditions with controlled changes in those conditions – interventions that may be evident, blind, or double blind – with statistical analysis to demonstrate linkages between the physical environment and outcomes. Evidenced-based design Use of laboratory- and field-gathered evidence in design decisionmaking. Evidenced-based design was adopted by the Center for Health Design to improve patient health and safety outcomes through improvements in design. LEED™, BREEAM™, Greenstar™, CASBEE™ Rigorous, voluntary sustainability certification systems developed in the USA, the UK, Australia, and Japan, respectively, that span land use and site, energy and atmosphere, water, materials and resources, and indoor environmental quality goals. Mixed-mode or hybrid HVAC An approach to space conditioning that combines natural ventilation from operable windows or vents with mechanical heating, cooling, and ventilation systems (HVAC). Precautionary principle Adopted by the European Community and several nations around the world, the precautionary principle argues that if an action or policy has a

# Springer Science+Business Media, LLC, part of Springer Nature 2020 V. Loftness (ed.), Sustainable Built Environments, https://doi.org/10.1007/978-1-0716-0684-1_197 Originally published in R. A. Meyers (ed.), Encyclopedia of Sustainability Science and Technology, # Springer Science+Business Media LLC 2018 https://doi.org/10.1007/978-1-4939-2493-6_197-3

87

88

suspected risk of causing harm to the public or to the environment, even in the absence of scientific consensus that the action or policy is harmful, the burden of proof that it is not harmful falls on those taking the action. Triple bottom line Expanding the criteria for evaluating a project’s success from economic benefits alone to include ecological and social cost benefits, adopted by the United Nations and others to ensure public sector decision making reflects full cost accounting, and colloquialized as “people, planet, and profit.” WELL™ and FITWEL™ Emerging voluntary certification systems for community and building design to ensure higher levels of human health, addressing a range of factors including air, water, nourishment, light, fitness, comfort, and mind, as well as safety, community health, and social equality.

Definition of the Subject While sustainable design is focused on reducing the environmental footprint, the resources consumed, and the waste produced, it is also critically linked to our health. Design decisionmaking for sustainability – land use, building massing and enclosure, lighting systems, mechanical systems, interior systems, building operation, and management – can not only reduce our environmental footprint; it can and must enhance our visual, aural, dermal, musculoskeletal, circulatory, respiratory, reproductive, and mental health. The challenge is to explore the linkages between critical design decisions, from land use to material and system design to building maintenance and operations, to include critical health outcomes. Based on years of gathering emerging laboratory, field, and epidemiological case studies, the Center for Building Performance and Diagnostics at Carnegie Mellon University has been assembling the research on environmental, health, and productivity benefits of the range of high-performance building systems that are the basis for sustainable design. Captured in a Building Investment Decision Support tool (BIDS™), the cost benefits of investing

Sustainable and Healthy Built Environment

in a sustainable built environment revealed in these case studies should drive measurable changes in building design, construction, and management. Given critical data sets linking sustainable design solutions and health, triple bottom line accounting can support strategic, longterm decision-making in the built environment, commensurate with its intended life of 30–50 to hundreds of years. At the same time, the funding for ongoing research and the aggregation of knowledge linking health and the built environment within an international data base will be critical to ensuring the necessary investment in the quality of our buildings and communities – the foundation of sustainability.

Introduction: Linking Definitions of Sustainability and Health Many decision-makers assume that sustainable design is about resource conservation – energy, water, and material resources. The last 10 years, however, has seen a dramatic broadening of the definition of sustainability to include commitments to mobility through improvements in land use and transportation, commitments to health and productivity through improvements in indoor environmental quality and active design, and commitments to the protection of regional strengths and a more globally shared quality of life. In the USA, this broader definition of sustainability is most often ensured through the LEED™ standard (Leadership in Energy and Environmental Design) of the US Green Building Council. Paralleled by developments in the UK (BREEAM), Australia (Greenstar), and Japan (CASBEE), the 100 or more credits, these sustainability rating systems for new and existing buildings extend beyond conservation goals for energy, water, and materials, to include sustainable sites and transportation, indoor environmental quality, and regional sustainability (usgbc.org/LEED). In the past 5 years, a number of standards focused on the design of built environments for human health have emerged, including the WELL™ Standard from the International Well Building Institute (IWBI) and the FITWEL™

Sustainable and Healthy Built Environment

Standard from the Center for Active Design, the US GSA, and the US Center for Disease Control. The WELL standard addresses seven health and wellness categories: air, water, nourishment, light, fitness, comfort, and mind. The FITWEL standard’s health impact categories are community health, well-being, physical activity, reduced morbidity and absenteeism, healthy foods, safety, and social equality for vulnerable populations. The Center for Building Performance and Diagnostics at Carnegie Mellon University would argue for expanding this definition even further, to give even greater emphasis to contextual and regional design goals, to accessible and flexible infrastructures that support change, to natural/passive conditioning, and to human engagement, motivation, and health. Indeed, the CBPD research team proposes the following definition: Sustainable design is the result of a transdisciplinary, integrated design process driven to ensure that the built environment achieves greater levels of ecological balance in new and retrofit construction, toward the long term viability and humanization of architecture. Focusing on environmental context, sustainable design merges the natural, minimum resource conditioning solutions of the past (daylight, solar heat and natural ventilation) with the innovative technologies of the present, into an integrated “intelligent” system that supports individual control with expert negotiation for environmental quality and resource consciousness. Sustainable design rediscovers the social, environmental and technical values of pedestrian, mixed-use communities, fully using existing infrastructures, including “main streets” and small town planning principles, and recapturing indooroutdoor relationships. Sustainable design avoids the further thinning out of land use, and the dislocated placement of buildings and functions caused by single-use zoning. Sustainable design introduces benign, non-polluting materials and assemblies with lower embodied and operating energy requirements, and higher durability and recyclability. Finally, sustainable design offers architecture of long term value through “forgiving” and modifiable building systems, through life-cycle instead of least-cost investments, and through timeless delight and craftsmanship. [1]

The depth of the definition of sustainability matters, especially when assessing the relevance of sustainable design, construction, and operations of buildings for long-term human and

89

environmental health. This chapter explores the range of health outcomes that can be linked to these important principles for the design of a sustainable built environment:

CMU: Seven Principles for the Design of Sustainable Built Environments 1. Sustainable design depends on an integrative, human-ecological design approach. 2. Sustainable design depends on changing approaches to land use and community fabric. 3. Sustainable design depends on the promotion of infrastructures to neighborhood amenities. 4. Sustainable design depends on the effective use of natural, local, and global resources to reduce resource demands and maximize resource use. 5. Sustainable design depends on the design of flexible, forgiving systems. 6. Sustainable design depends on the use of healthy, renewable materials and assemblies. 7. Sustainable design depends on design for life cycle instead of first cost.

A Definition of Health to Be Integral with Sustainable Design Equally critical to a shared definition of sustainable design is the exploration of a shared definition of human health. Building on the Cornell Medical Index of 1949 [2], the Center for Building Performance at Carnegie Mellon proposed ten health indices as a critical to design decisionmaking in 2010 (Fig. 1). In 2014, the International Well Building Institute launched the WELL Building Standard ® – a performance-based system for measuring, certifying, and monitoring features of the built environment that impact human health and well-being, through air, water, nourishment, light, fitness, comfort, and mind. Based on these precedents, in-depth consideration of at least 12 qualities of human health will be critical to valuating building design, construction, and operational decisions:

90

Sustainable and Healthy Built Environment

Sustainable and Healthy Built Environment, Fig. 1 Correlating decision-making in the built environment to improving human health should be a major international initiative (figure CBPD)

Sustainable and Healthy Built Environment

Qualities of human health integral with sustainable design: 1. Respiratory health 2. Digestive health 3. Visual health 4. Aural health 5. Skin/dermal health (integumentum system) 6. Musculoskeletal health 7. Cardiovascular health 8. Nervous system health 9. Genitourinary (including reproductive) health 10. Endocrine system health 11. Immune system health 12. Mental health

To date, evidence from research suggests six primary clusters of health issues related to the built environment: respiratory (chest, wheeze, allergies, asthma, colds, flu), mucosal (eye, nose, throat), dermal (face, hand skin), neurological (e.g., headache, migraine, dizziness, heavyheadedness), musculoskeletal (to include body mass today), and mental health. The challenge is to definitively link these health issues to the quality of the built environment. Given the unbounded variations in physical settings and their management over time, arguments for causality between a single design decision, such as type of ventilation system, and health outcome will be very difficult. Moreover, funding for multivariate research questions is scarce. The challenge is to identify the controlled experiments, field intervention studies, and portfolio-wide studies that might support effective design decisionmaking for sustainability. The Center for Health Design (http://www.healthdesign.org/) has made the case for evidenced-based design that “is the process of basing decisions about the built environment on credible research to achieve the best possible outcomes.” They have launched the “Pebble Project” to capture profiles of healthcare organizations whose facility design has made a difference in the quality of care as well as their financial performance. These case studies explore the potential benefits of daylight and view, acoustics, variations in ventilation, hospital room layout

91

and configuration, and more, on hospital outcomes. These projects are changing design practice, but some might argue they are not based on definitive research that can explain the mechanisms whereby health outcomes are improved. For this to occur, multiyear epidemiological studies as well as controlled laboratory and field experiments must be undertaken by collaborative teams of health and building professionals. In the meantime, the design community is making decisions every day that can positively or negatively impact human health. To this community, it is argued that every study available must contribute to informed decision-making and every act must be based on the European Union’s precautionary principle: Where there is uncertainty as to the existence or extent of risks of serious or irreversible damage to the environment, or injury to human health, adequate protective measures must be taken without having to wait until the reality and seriousness of those risks become fully apparent. (http://www.sustainabledesign.ie/arch/precautionaryprinciple.htm)

The precautionary principle argues for the building community to take responsibility to protect the public from exposure to harm, whenever scientific investigation has found a plausible risk. These protections can be relaxed only if further scientific findings emerge that provide sound evidence that no harm will result. For example, scientific investigation suggests that outdoor air ventilation rates should be doubled in buildings to reduce the risks of colds, flus, and respiratory illnesses. Research further argues that these increased ventilation rates should be accompanied by economizer cycles and/or heat recovery to ensure that primary energy loads will not increase, since power plant pollution also poses measurable health risks.

Linking Health and the Built Environment By setting a definition of the attributes of sustainable design against the characteristics of human health, even intuitive judgment would illuminate the importance of building design, construction, and operation for human health (Fig. 1).

92

With over 10 years of intense study by faculty, researchers, and graduate students, the Center for Building Performance and Diagnostics at Carnegie Mellon has been collecting building case studies as well as laboratory and simulation studies in an effort to statistically link the quality of buildings – system by system – to productivity, health, and life cycle sustainability. Amassed in the BIDS™ (Building Investment Decision Support) tool, these case studies enable building decision-makers to calculate return on investments for high-performance building components and systems and will lead to greater understandings of the importance of buildings and communities to human health (see http://cbpd. arc.cmu.edu/bidstrial). The following six sections explore design innovations and the potential health impacts of changes in land use, building massing and enclosure, HVAC engineering, daylight and lighting system design, interior systems, and longterm building maintenance and operations.

Sustainable Land Use for Health One of the most significant design shifts needed for the long-term health of humans is to move away from the automobile-centric land use planning that makes pedestrian lifestyles impossible. The dramatic reductions in walking and biking that have resulted from sprawl and single-use zoning have contributed to increasing rates of obesity in industrialized nations [3–6]. At the same time, the increased reliance upon automobiles has resulted in ever-increasing levels of particulate and ozone that are respiratory and cardiovascular hazards [7]. David L. Skole of the Center for Global Change and Earth Observations at Michigan State University identifies a range of health effects from sprawl: air pollution, CO2 emissions, heat island effect, reduced physical exercise, increased car accidents, and pedestrian injuries, as well as declining water quality [8]. Numerous studies have revealed the seriousness of particulate-related health concerns. Wordley et al. [9] identified a 2.4% increase in respiratory hospital admissions and a 2.1% increase in cerebrovascular admissions associated with a 10 mg/m3 increase in PM10 in the air, which, according to Dockery &

Sustainable and Healthy Built Environment

Pope [10], increases respiratory admissions by 0.8  3.4%. Tenias et al. [11] found that a 10 mg/ m3 increase of NO2 and O3 in the air causes increases in the number of emergency visits for asthma by 7.6% and 6.3%, respectively. Moreover, automobile-centric land use and single-use zoning has led to increasing inactivity in children, teens, and senior citizens that contributes to obesity and all the associated health impacts of obesity, as well as potentially contributing to attention deficit, depression, and suicide. In a 2004 cross-sectional survey of 10,878 adults from the Atlanta, Georgia region, Frank et al. identify a 12.2% decline in the likelihood of obesity for each quartile increase in land use mix, which was defined on a scale from 0 (indicating a single-use environment) to 1 (indicating a perfectly equal mix of residential, commercial, office, and institutional uses). They also found a significant relationship between reductions in vehicle miles travel and reductions in particulate matter, NOX and VOC pollution, which have other health ramifications (see Fig. 2), [12]. The US studies on land use and obesity are summarized and advanced in the 2006 National Institute of Health study of the “Relationship Between Urban Sprawl and Weight of United States Youth” [13]. “The first studies reporting a direct relationship between the built environment and obesity were published in 2003 [3, 14, 5, 6]. After controlling for age, education, fruit and vegetable consumption, and other socio-demographic and behavioral covariates, Ewing et al. [4] found that adults living in sprawling counties had higher body mass indices (BMIs) and were more likely to be obese (BMI  30) than were their counterparts living in compact counties. Independent studies have since generally confirmed these original findings. Specifically, all macro-level (county or larger) studies, and all but one micro-level (neighborhood) studies, have found significant relationships, in the expected direction, between sprawl-like development patterns and BMI, after controlling for sociodemographic and other influences” [13]. Moreover, automobile-based design is “paving” the countryside, with the elimination of landscapes that act as natural lungs for filtering our air, and natural digestive systems for processing the increasing levels of salts, oils, and storm sewer

Sustainable and Healthy Built Environment

93

Increased density + mixed land use = Health

Land use mix

Frank et al 2004

Black male

White female

Black female

0.40

Probability of obesity

In a 2004 cross-sectional survey of 10,878 adults from the atlanta, Georgia region, Frank et al identify a 12.2% decline in the likelihood of obesity for each quartile increase in land use mix, which was defined on a scale from 0 (indicating a single-use environment) to 1 (indicating a perfectly equal mix of residential, commercial, office, and institutional uses).

White male

Average mixed use, Atlanda

0.30

Maximum mixed use, Atlanda

0.20

0.10

0.00 0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00 8.00 9.00 10.00

Walk distance per day in kilometers White male

Black male

White female

Black female

First cost saving: Annual health savings: Rol:

$4,408 / person $51 / person Immediate

Probability of obesity

0.40

0.30

0.20

0.10

0.00 0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00 8.00 9.00 10.00

Reference: D. Frank, M. Andresen, T. Schmid. Obesity relationships with community design, physical activity, and time spent in cars. American Journal of Preventive Medicine, Volume 27, Issue 2, August 2004. Center for Building Performance and Diagnostics, a NSF/IUCRC, and ABSIC at Carnegie Mellon

Sustainable and Healthy Built Environment, Fig. 2 Land use can significantly reduce vehicle miles traveled which in turn reduces the pollution that causes growing respiratory health concerns

overflows that result in toxic runoff into our drinking water. The quantification of these serious health hazards should fully justify the shift in sustainable design to live-work-walk lifestyles with mixed-use communities, multigenerational mobility with mixed-mode transportation, and the preservation and celebration of natural landscapes and sustainable infrastructures. Linking sustainable land use guidelines with health • Design live-work-walk communities to reduce car pollution – particulates and ozone – that trigger asthma • Design for pedestrian, bicycle, transit mobility to reduce obesity • Minimize paving for roads and parking to reduce salting and oil runoff, as well as standing water concerns (continued)

Linking sustainable land use guidelines with health • Design landscape dominant environments to reduce thermal heat islands and heat stress and to rebuild nature’s lungs for air quality

Sustainable Building Massing and Enclosure for Health After land use design, the second most critical design decision for human health might be building massing and enclosure specifications. On the one hand, access to nature’s assets – daylight, natural ventilation, natural comfort and thermal variation, views, and physical access to outdoor

94

activity – is becoming increasingly linked to human health. On the other hand, humans need protection from nature’s liabilities – overheating, excessive cold, wind, rain, and snow. The design of the building enclosure is critical for managing both these climatic assets and liabilities. The CBPD has identified a range of international case studies linking access to the natural environment – daylight, views, and natural ventilation – to improved health outcomes, including reductions in headaches, colds, sick building symptoms (SBS), and patient length of stay (see Fig. 3). Beyond the health benefits, a number of international case studies demonstrate that access to the natural environment increases individual productivity between 3% and 18%, reduces absenteeism between 9% and 71%, and offers over 50% lighting and HVAC energy savings (CBPD). While the debate continues as to the mechanisms by which daylight improves health, research continues to reveal that sunlight, especially morning sunlight, reduces length of stay for patients recovering from surgery, bipolar, and SAD treatment [15–18]. In a 2005 building case study of Inha University Hospital in Korea, Choi identifies a 41% reduction in average length of stay among gynecology patients in bright (sunlit) rooms, as compared to those in dull rooms, in spring, and an average 26% reduction in average length of stay among surgery ward patients in bright rooms, as compared to those in dull rooms, during spring and fall [17]. In a separate study of pain medication use among 89 patients undergoing elective cervical and lumbar spinal surgery at Montefiore Hospital in Pittsburgh, PA, Walch et al. identify a 22% reduction in analgesic medication use among patients in bright rooms who were exposed to more natural sunlight after surgery, as compared to patients located in dim rooms after surgery [18] (Fig. 4). The work of the Lighting Research Institute at RPI has begun to reveal the possible mechanisms of these health outcomes, identifying the relationship of ultraviolet light exposure to the production of melatonin, a natural hormone that controls circadian rhythms that are related to sleep cycles and potentially to reduced cancer cell development

Sustainable and Healthy Built Environment

[19]. At the same time, daylight and sunlight can introduce glare and overheating that might negatively affect human health outcomes. The importance of views of nature and proximity to windows with views may be even more important for human health. The work of Ulrich [20], Mendell [21], Heschong Mahone Group [22], and Kellert [23] identify links between views and reduced length of hospital stay after surgery, reduced sick building syndrome, improved emotional health, and improved performance at task. In a 1991 11-building study of office buildings in San Francisco, Mendell [21] identifies 25–52% reductions in reported SBS symptoms among occupants located within 15 ft. of a window, as compared to those seated further from a window. In addition to confirming the importance of seated views for all building occupants, research is critically needed to understand the importance of the content of those views from windows for human health (e.g., the benefit of landscape views over parking lots, building walls, and sky). In a seminal 1984 field study at a Pennsylvania hospital, Ulrich [20] identifies an 8.5% reduction in postoperative hospital stay (7.96 vs. 8.7 days) for gall bladder surgery patients who had a view of a natural scene from their hospital room, as compared to those with a view of the brick wall of the adjacent building wing. Patients with a view of nature also received fewer negative evaluations from nurses and took fewer strong analgesics (see Fig. 5). In addition to sunlight and views, it is critical to understand the benefits of direct access to outdoor air and outdoor spaces through operable windows and doors. The value of increasing outside air delivery rates is becoming increasingly evident, as will be described in the section on HVAC design. It is not clear, however, whether increased levels of outside air are more effectively delivered through natural ventilation (operable windows) or through mechanical systems that incorporate filtration, dehumidification, and thermal conditioning of that outside air. There are over a dozen studies that reveal the benefits of natural ventilation in existing buildings as compared to mechanically ventilated buildings – benefits that

Sustainable and Healthy Built Environment, Fig. 3 A range of international case studies link access to the natural environment – daylight, views, and natural ventilation – to improved health outcomes, reductions in headaches, colds, sick building symptoms (SBS), and patient length of stay

Sustainable and Healthy Built Environment 95

96

Sustainable and Healthy Built Environment

Daylight = Health Average length of stay in bright and dull rooms in two units

Inha University Hospital / Choi 2005 (Hospital)

First cost increase: Annual health savings: ROI:

10 Average length of stay (days)

In a 2005 building case study of Inha University Hospital in Korea, Choi identifies a 41% reduction in average length of stay among gynecology patients in bright (sunlit) rooms, as compared to those in dull rooms, in spring, and an average 26% reduction in average length of stay among surgery ward patients in bright rooms, as compared to those in dull rooms, during spring and fall. $1,000 / bed $10,115 / bed 1,011%

Bright rooms Dull rooms 8

6

4

2

0

Gynecology (spring)

Surgery (fall)

Reference: Choi, Joonho. (2005). Study of the Relationship between Indoor daylight Environments and Patient Average Length of Stay (ALOS) in Healthcare Facilities, Unpublished master’s thesis, De[artment of Architecture, Texas A&M University. College Station, TX. Center for Building Performance and Diagnostics, a NSF/IUCRC, and ABSIC at Carnegie Mellon

Daylight = Health Montefiore Hospital / Walch et al 2005 (Hospital)

First cost increase: Annual health savings: ROI:

$1,000 / bed $28 / bed 3%

8 Mean oral morphine consumption (mg/hr)

In a 2005 study of pain medication use among 89 patients undergoing elective cervical and lumbar spinal surgery at Montefiore Hospital in Pittsburgh, PA, Walch et al identify a 22% reduction in analgesic medication use among patients in bright rooms who were exposed to more natural sunlight after surgery, as compared to patients located in dim rooms after surgery.

Average medication use per day by room type

Bright room Dim room

7 6 5 4 3 2 1 0

Surgery Post-op Post-op Post-op Post-op Post-op day day 1 day 2 day 3 day 4 day 5

Reference: Walch JM, Rabin BS, Day R, Williams JN, Choi K, Kang JD (2005) The effect of sunlight on postoperative analgesic medication use: a prospective study of patients undergoing spinal surgery. Psychosom Med 67:156–163 Center for Building Performance and Diagnostics, a NSF/IUCRC, and ABSIC at Carnegie Mellon

Sustainable and Healthy Built Environment, Fig. 4 In extensive hospital record studies, Choi and Walch identified the value of southern and southeastern rooms to reduced patient length of stay

range from reduced headaches, mucosal symptoms, colds, coughs, and circulatory problems, to reduced SBS symptoms. In a 1992 study of two London hospitals, Kelland [24] identifies a 40%

reduction in sick building syndrome (SBS) symptoms among staff of a naturally ventilated hospital, as compared to those of a mechanically ventilated hospital (see Fig. 6). In a 2004 multiple

Sustainable and Healthy Built Environment

97

Seated Access to Views = Health

Floor plan at a hospital Corridor

PA Hospital / Ulrich 1984

Patients’ rooms

First cost increase: Annual health savings: ROI:

$ 5,000 /bed $ 2,237 / bed 45%

Brick wall

Recovery room

Average length of stay with a view of nature vs without a view of nature Average length of stay (days)

In a 1984 observational field study at a Pennsylvania hospital, Ulrich identifies an 8.5% reduction in post-operative hospital stay (8.7 vs 7.96 days) for gall bladder surgery patients who had a view of a natural scene from their hospital room, as compared to those with a view of a brick wall. Patients with a view of nature also received fewer negative evaluations from nurses and took fewer strong analgesics.

10 8 6 4 2 0 Room w/a view of nature Room w/a view of brick wall

Reference: Ulrich, R. (1984) View Through a Window May Influence Recovery From Surgery. Science, 224 (4647), pp. 420-421.

Center for Building Performance and Diagnostics, a NSF/IUCRC, and ABSIC at Camegie Mellon

Seated Access to Windows = Health + Individual Productivity CA Healthy Building Study / Mendell 1991 (Wargocki et al 2000) In a 1991 multiple building study of 11 office buildings in San Francisco, Mendell identifies 25% - 52% reductions in reported SBS symptoms among occupants located within 15 feet of a window, as compared to those seated further from a window In a 2000 study, Wargocki et al identify a 1.1% productivity increase for every 10% reduction in SBS complaints, suggesting an average 4.3% productivity gain for workers seated near a window. First cost increase: Annual productivity savings: Annual health savings: ROI:

$1,000 / employee $1,935 / employee $40 198%

Reported symptoms by distance to window

Window < 15 ft Window < 15 ft

Headache

Chills or fever

Eye, nose, throat symptoms 0

20

40

60

% reporting symptoms

Reference: Mendell, Mark J. (1991) Risk Factors for work-Related Symptoms in Northern California Office Workers. Wargocki, P,Wyon, D. and Fanger, P.O. (2000) Pollution Source Control and Ventilation Improve Health, Comfort and Productivity. In Proceedings of Cold Climate HVAC 2000, Sapporo, Japan, November 1-3, 2000, pp. 445-450.

Center for Building Performance and Diagnostics, a NSF/IUCRC, and ABSIC at Carnegie Mellon

Sustainable and Healthy Built Environment, Fig. 5 Ulrich and Mendell identified the importance of views and view content to sick building syndrome and reduced patient length of stay

building study of professional middle-aged women in France, Preziosi et al. [25] identify a 57.1% reduction in sickness absence, a 16.7%

reduction in medical services use (doctor visits), and a 34.8% reduction in hospital stays among subjects with natural ventilation in their

98

Sustainable and Healthy Built Environment

Natural Ventilation = Health + Individual Productivity Per capita health services use and absenteeism in naturally ventilated vs. mechanically ventilated buildings

Preziosi et al 2004 In a 2004 multiple building study of professional middle-aged women in France, Preziosi et al identify a 57.1% reduction in sickness absence, a 16.7% reduction in medical services use (doctor visits), and a 34.8% reduction in hospital stays among subjects with natural ventilation in their workplace, as compared to those with air conditioning.

4.5 Natural ventilation Air conditioning

4 3.5 3 2.5 2 1.5 1

First cost increase: Annual health savings: Annual productivity savings: ROI:

$1,000 / employee $181 / employee $85 / employee 27%

0.5 0 Doctor’s visits (number)

Hospital stays (days)

Sickness absence (days)

Reference: Preziosi P., S.Czernilchow, P.Gehanne, and S. Hercberg (2004) Workplace air-conditioning and health services attendance among French middle-aged women: a prospective cohort study. International Journal of Epidemiology, 33(5). pp. 1120-1123.

Center for Building Performance and Diagnostics, a NSF/IUCRC, and ABSIC at Carnegie Mellon

Natural Ventilation = Health London Hospitals / Kelland 1992 (Wargocki et al 2000)

In a 2000 study, Wargocki et al identify a 1.1% increase in productivity for every 10% reduction in SBS complaints, suggesting a 4.4% productivity gain due to natural ventilation. First cost increase: Annual productivity savings: ROI:

$543 / employee $1,980 / employee 365%

SBS symptom prevalence at a naturally ventilated and a mechanically ventilated hospital

45% RFH (mech, vent) MH (net vent)

40% % reporting symptom

In a 1992 multiple building study of two London hospitals, Kelland identifies a 40% reduction in sick building syndrome (SBS) symptoms among staff of a naturally ventilated hospital, as compared to those of a mechanically ventilated hospital.

35% 30% 25% 20% 15% 10% 5% 0% ry

D

at

in

sk

re

So

ro

th

H

he

ac

d ea

S

e or

es rgy ess ose ose sea ey tha n si y n y n Nau Le row nn tuff D Ru S

Reference: Kelland, P (1992) Sick Building Syndrome, Working Environments and Hospital Staff. Indoor Environment, v1, PP.335-340; Wargocki, P, Wyon, D, and Fanger, P.O. (2000) Pollution Source Control and Ventilation Improve Health, Comfort and Productivity. In Proceedings of Cold Climate HVAC 2000, Sapporo, Japan, November 1-3 2000, pp. 445-450. Center for Building Performance and Diagnostics, a NSF/IUCRC, and ABSIC at Carnegie Mellon

Sustainable and Healthy Built Environment, Fig. 6 Kelland identified a reduction in hospital staff SBS, and Preziosi identified reductions in absenteeism

and doctor visits in buildings with natural ventilation as compared to those with mechanical ventilation or air conditioning

Sustainable and Healthy Built Environment

workplace, as compared to those with air conditioning only (see Fig. 6). In a 2007 study of 104 child care centers in Singapore, Zuraimi et al. identify a 10.7% decrease in cough with cold/flu symptoms, a 31% decrease in phlegm with cold/flu, a 27.3% decrease in coughing attacks lasting more than 1 week, and a 33.3% decrease in lower respiratory illness in children attending naturally ventilated child care centers as opposed to centers with hybrid ventilation [26]. While operable windows can bring in higher quantities of outside air, they can also bring in unwanted outdoor pollution, humidity, rain, and noise. The pros and cons of increasing outside air rates through natural versus mechanical means are outlined in Fig. 3, with a nod to the value of natural ventilation, especially given the variable long-term performance of HVAC systems and controls in operation (Fig. 7). The design decisions central to ensuring daylight, view, and natural ventilation include increasing surface area with thinner floor plates and resolving glare, overheating, heat loss, and rain penetration through appropriate enclosure design. In some respects, sustainable, healthy buildings have many of the characteristics of sustainable, healthy humans – they are physically fit rather than obese (thin floor plans, finger plans, and courtyard buildings); they have circulatory systems that take the excess heat from the core out to the surface (through water mullions or air flow windows for example); and they absorb sunlight and breathe fresh air. At the same time, sustainable buildings are designed to reduce or make effective use of climatic liabilities – rain, cold and hot temperatures, diurnal temperature

99

swings, excessive sun, and freeze-thaw – with completely regional design solutions. Linking sustainable building massing and enclosure guidelines with health • Design for daylighting without glare to support visual acuity and reduce headaches • Design for natural ventilation without drafts and rain penetration to reduce respiratory symptoms • Engineer thermal load balancing to eliminate radiant asymmetry associated with arthritis and circulatory disorders • Design for passive solar heating where climate appropriate for thermal comfort and UV benefits • Design enclosure integrity to eliminate mold affecting SBS, respiratory/allergy, and asthma

Sustainable HVAC for Health The design of heating, ventilation, and air conditioning systems (HVAC) for human health is based on at least three improvements in individual occupant conditions: increased outside air rates and filtration, improved moisture/humidity control, and improved thermal comfort control. In addition to managing sources of pollution, healthy indoor air is dependent on a commitment to improving the quality and quantity of outside air. Increasing outside ventilation rates for health has substantial research justification – a doubling or tripling of code requirements for outside air measurably reduces headaches, colds, flus, nasal symptoms, coughs, and SBS symptoms [27–36]. Increasing outside air rates without energy penalty may be achieved by maximizing natural ventilation; by mixed-mode HVAC systems that support natural ventilation; by designing

Sustainable and Healthy Built Environment, Fig. 7 Operable windows are critical to long-term sustainability and human health, so challenges should be addressed through design

100

HVAC systems with separate ventilation air and thermal conditioning systems, which permits thermal conditioning to be water or air based; or by increasing outside air quantities with effective filtration and heat recovery for energy efficiency. In a meta-analysis of 20 ventilation rate studies, Seppanen, Fisk, and Mendell found that the relative risks of respiratory illnesses were 1.5–2 times higher in low ventilation rate buildings (below 10 L/s/person) as compared to high ventilation rate buildings (up to 20 L/s/person), alongside a 1.1–6 higher risk of sick building syndrome symptoms in low ventilation rate buildings [37]. A recent international study of papers published in peer-reviewed scientific journals identified a “biological plausibility for an association of health outcomes with ventilation rates” [38]. Higher ventilation rates in offices, up to about 25 L/s per person, are associated with reduced prevalence of sick building syndrome (SBS) symptoms. The limited available data also suggest that inflammation, respiratory infections, asthma symptoms, and short-term sick leave decrease with higher ventilation rates. Home ventilation rates above 0.5 air changes per hour (h 1) have been associated with a reduced risk of allergic manifestations among children in a Nordic climate. However, the literature does not provide clear evidence on particular agent(s) for the effects. The need remains for more studies of the relationship between ventilation rates and health impacts, especially in diverse climates, in locations with polluted outdoor air, and in buildings other than offices. To ensure ventilation effectiveness, the ventilation system must be designed to provide air to the individual with “task” air systems, ideally with some level of individual control to address local pollutant buildup. In a 2002 controlled experiment, Kaczmarczyk et al. identify a 23.5% reduction in headache symptoms when workers are provided with individually controlled task air systems supplying outdoor air, as compared to a conventional mixing ventilation system, in a room with a typical office pollutant source [39] (see Fig. 8). In a building intervention study in 39 Swedish schools, Smedje and Norback identify a 69% reduction in the 2-year incidence of asthma among students in schools that received a new displacement ventilation system with increased fresh air supply rates, as compared

Sustainable and Healthy Built Environment

to students in schools that did not receive a new ventilation system (see Fig. 8, [34]. Both of these studies reveal the health benefits of ensuring that fresh air reaches the nose of every occupant. At the same time, a healthy HVAC system must guarantee outdoor and indoor pollution source control through better building design, interior design, as well as HVAC configuration, filtration, and maintenance. In a 2003 multiple building study of three Montreal office buildings, Menzies et al. identify a 25% reduction in reported mucosal symptoms and a 25% reduction in reported respiratory symptoms due to ultraviolet germicidal irradiation (UVGI) of drip pans and cooling coils within ventilation systems [40]. Additional studies are critically needed to address the importance of both filtration and ongoing HVAC maintenance for human health. The CBPD has summarized a range of international case studies that link high-performance ventilation strategies to 10–90% reductions in respiratory illnesses, including asthma and allergies, as well as studies that demonstrate reductions in SBS, headaches, flus, and colds (see Fig. 9). In these studies, the critical HVAC improvements are increasing outside air rates, ensuring mold/moisture control and air stream management and maintenance, as well as ensuring quality filtration. In addition to health benefits, the CBPD has identified studies that link individual productivity gains of 1.7–11% to highperformance ventilation strategies. Studies also reveal that the energy penalty for increasing outside air rates can be easily eliminated with heat recovery or the use of mixed-mode ventilation and conditioning, to generate up to 50–80% energy savings in sustainable HVAC systems (CBPD). Sustainable HVAC systems must also be designed to provide individual thermal controls. Several laboratory and field experiments link temperature control to individual productivity gains between 0.2% and 7% to 15% energy savings through task thermal conditioning (CBPD) and to reduced headache and SBS symptoms [39, 41]. The challenges for HVAC design for thermal comfort are to design for dynamic thermal zone sizes (anticipating changing density and uses), provide individual thermal controls (separating thermal and air delivery);, design for building load balancing and

Sustainable and Healthy Built Environment

101

Floor-based ventilation + Increased outside air = Health Smedje & Norback 2000 (School) Two-year incidence of symptoms in students attending schools with and without new ventilation systems 14 New ventilation system 12 No New ventilation system

In a 2000 multiple building study of 39 schools in Sweden, Smedje and Norback identify a 69% reduction in the 2-year incidence of asthma among students in schools that received a new displacement ventilation system with increased fresh air supply rates, as compared to students in schools that did not receive a new ventilation system.

10 8 6 4 2

First cost increase: Annual energy cost increase: Annual health savings: ROI:

$38 / student $2 / student $36 / student 89%

0

Pollen/pet Asthma ever allergy

Current asthma

Any asthma More asthma symptoms symptoms in 1995 than 1993

Reference: Smedje. G and Norback, D.(2000) New ventilation systems at select schools in Sweden–Ettects on Asthma and Exposure. Archives of Environmental Health, 35(1), pp. 18-25.

Center for Building Performance and Diagnostics, a NSF/IUCRC. and ABSIC at Carnegie Mellon

Individual air control + Increased outside air = Health Kaczmarczyk et al 2002

In a 2002 controlled experiment, Kaczmarczyk et al identify a 23.5% reduction in headache symptoms when workers are provided with individually-controlled task air systems supplying outdoor air, as compared to a conventional mixing ventilation system, in a room with a typical office pollutant source. First cost increase: Annual energy cost increase: Annual health savings: Annual productivity savings: ROI:

Reported headache symptoms by type of ventilation system

Severe headache 20

$800 / employee $8 / employee $17 / employee $106 / employee 14%

18 16 14 12 10 8 6 4 2 No headache 0

Task air, 20;C fresh

Task air, 23;C fresh

Task air, 23;C recirculated

Mixing ventilation, 23;C

Reference: Kaczmarczk, J., Zeng. Q., Melikov. A., and Fanger. P.O. (2002) The effect of a personalized ventilation system on perceived air quality and SBS symptoms. In Proceedings of Indoor Air 2002, Monterey. CA. June 30-July 5, 2002; Office of Pollution Prevention and Toxics, U.S. Environmental Protection Agency. Cost of Illness Handbook. http://www.epa. gov/oppt/coi; Schwartz et al (1997) Lost Workdays and Reduced Work Effectiveness Associated with Headache in the Workplace. Journal of Occupational and Environmental Medicine. 39(4). pp. 320-327.

Center for Building Performance and Diagnostics, a NSF/IUCRC, and ABSIC at Carnegie Mellon

Sustainable and Healthy Built Environment, Fig. 8 Smedje and Norback [34] identify a 69% reduction in 2-year incidences of asthma in schoolchildren, and

Kaczmarczyk [39] identifies a 23.5% reduction in headache symptoms in offices when ventilation is delivered more effectively to individuals

Sustainable and Healthy Built Environment, Fig. 9 A range of international case studies link improvements in HVAC to reduce colds, headaches, respiratory, mucosal, and SBS symptoms

102 Sustainable and Healthy Built Environment

Sustainable and Healthy Built Environment

radiant comfort, and, finally, engineer and prototype robust systems that provide air quality and thermal comfort consistently in the field, over time. Linking sustainable HVAC guidelines with health • Increase outside air rates, through natural ventilation or HVAC systems with heat recovery – To reduce respiratory, allergy, asthma, colds, headaches, and SBS • Engineer ventilation effectiveness, including air path and filtration management – To reduce respiratory, throat, and mucosal symptoms • Engineer moisture/humidity management – To reduce mold affecting respiratory illnesses, colds, and SBS • Separate ventilation and thermal conditioning systems for individual thermal control – To reduce headaches and SBS symptoms and to support local temperature and air quality control

Sustainable Lighting for Health Sustainable design must maximize the use of daylight for both sustainability and health, as long as it can be provided without glare and excessive heat loss or heat gain. Daylight can provide the higher light levels needed for fine work, improve color rendition and sculptural definition, give full spectrum and ultraviolet content that might be critical to circadian rhythms, and provide access to views of nature. Subsequently, electric lighting systems have the responsibility to effectively interface with daylight to meet the needs of specific tasks and provide the appropriate quantity and quality of light when daylight is not available. To this end, sustainable lighting is dependent on selecting the highest-quality lighting fixtures, lamps, ballasts, reflectors, lenses, and controls to light each specific task or task surface in an energy-effective manner. The benefits of well-designed daylight for human health have been previously discussed. If daylight is well designed to control glare and brightness contrast, it is a low-energy source that can significantly improve task performance and reduce headaches. In addition, the spectral distribution of daylight, critical to plant health, as well as time-of-day variations in light, may have a measurable impact on our circadian rhythms that impact sleep cycles and energy levels. Finally, as previously discussed, views that may be associated with daylight sources may have a measurable

103

impact on reducing depression and SBS symptoms, while improving hospital recovery rates. The high-quality lighting fixtures, lamps, ballasts, reflectors, lenses, and controls to light each specific task or task surface also have measurable benefits for human health. For example, in a 1989 controlled field experiment at a government legal office in the UK, Wilkins et al. identify a 74% reduction in the incidence of headaches among office workers when magnetic ballasts are replaced by high-frequency electronic ballasts (see Fig. 10) [42]. Given the shift from paper-based to computerbased tasks, lighting design must be improved for task performance, for energy effectiveness, and for human health. Sustainable lighting design supports the separation of task and ambient lighting – to enable lower overall ambient light levels at 200–300 lux for computer-based work and face-toface discussions and to be augmented by higher task light levels at 500–800 lux for fine print work. In a 1998 multiple building study in Germany, Çakir and Çakir identify a 19% reduction in headaches for workers with separate task and ambient lighting, as compared to workers with ceiling-only combined task and ambient lighting (see Fig. 10) [43]. Three international case studies demonstrate that improved lighting design reduces headache symptoms, as shown in Fig. 11. The CBPD has also identified 12 international case studies that indicate that improved lighting design increases individual productivity between 0.7% and 23% while reducing annual energy loads by 27–88% (CBPD). Linking sustainable lighting guidelines with health • Design for daylighting without glare to support visual acuity, color rendition, circadian rhythms, and view content to reduce headaches and hospital length of stay • Specify high-performance fixtures for maximum lumens/watt, reduced glare, shadowing, and noise and to reduce headaches • Separate ambient and task lighting delivery to match light levels to task and provide individual control

Sustainable Interior Systems for Health: Materials and Ergonomics Among a range of interior design decisions that affect both sustainability and productivity, at least

104

Sustainable and Healthy Built Environment

High Performance Luminaires = Health Wilkins et al 1989 In a 1989 controlled filed experiment at a government legal office in the UK, Wilkins et al identify a 74% reduction in the incidence of headaches among office workers when magnetic ballasts are replaced by high frequency electronic ballasts.

40 reporting symptom per week

First cost increase: Annual health savings: Annual productivity savings from reduced headaches: ROI:

Frequecny of headaches among employees

$109 / employee $54 / employee

30

20

10

0

$333 / employee 355%

Electronic ballasts

Magnetic ballasts

Referecne: Wilkins, AJ, Nimmo-smith, I, Slater, AI, Bedocs, L. (1989)Fluorescent lighting, headches and eyestrain. Lighting Research and Technology 21(1), PP. 300-307. Center for Building Performance and Diagnostics, a NSF/IUCRC, and ABSIC at Carnegie Mellon

Lighting control = Individual productivity + Health Influence of lighting on the degree of disturbance to health (1=no disturbance, 4=strong disturbacne)

Cakir and Cakir 1988

First cost increase: Annual health savings: Annual productiviy savings: ROI:

$314 / employee $14 / employee $87 / employee 32%

4 Degree of Impairment

In a 1988 multiple building study in Germany, Cakir and Cakir identify a 19% reduction in headaches for workers with separate task and ambient lighting, as compared to workers with ceiling-only combined task and ambient lighting.

3

2

1 Table lamp

Overhead

Ceiling

Reference: Cakir, A.E. and Cakir, G. (1998) Light and Health: Influences of Lighting on Health and Well-being of Office and Computer Workers, Ergonomic, Berlin Center for Building Performance and Diagnostics, a NSF/IUCRC, and ABSIC at Carnegie Mellon

Sustainable and Healthy Built Environment, Fig. 10 Wilkins identified a 74% reduction in headaches with updated lighting ballasts, and Cakir and Cakir

identified 19% reduction in headaches with the separation of ambient and task lighting

three design decisions also have measurable health impacts – healthy material selection, acoustic/noise management, and the ergonomics of furniture and space layout. Interior material selection is critical to thermal performance, air quality (due to outgassing), toxicity in fires, cancer-causing fibers, and mold growth, which in turn impact our respiratory and digestive

systems, our eyes and skin. The CBPD has identified six studies linking materials selection to health outcomes including SBS, mucosal irritation, allergies, and asthma (see Fig. 12, [44–49]. In a 1996 study of 80 homes in Victoria, Australia, Garrett et al. identify a 60% reduction in the prevalence of asthma and a 63% reduction in the prevalence of allergies among children whose homes contain formaldehyde-free

Sustainable and Healthy Built Environment

Health Benefits of High Performance Lighting

100% 90%

CBPD ABSIC BIDSTM

Sustainable and Healthy Built Environment, Fig. 11 Case studies link improvements in lighting – ballast quality and individual control – to reduced headache symptoms

105

fewer headaches

80%

74%

70% 60% 50% 40% 27%

30%

19%

20% 10% 0%

user controlled indirect/direct fixtures Aaras et al 1998 [46]

high frequency electronic ballasts

split task and ambient lighitng

Wilkins et al 1989 [47] Cakir & Cakir 1998 [48]

Health Benefits of High Quality Interior Materials CBPD ABSIC BIDSTM

100% fewer SBS symptoms 90% 85% 80% fewer allergies

reduced asthma

70% 63% 60% 60%

50%

47%

40% 23% fewer mucosal symptoms

30% 24% 20% 20% 14% 10%

Jaakkola et al 1994 [52]

Jaakkola et al 1999 [55]

Wieslander et al 1997 [54]

Garrett et al 1996 [53]

Garrett et al 1996 [53]

Jaakkola et al 1994 [52]

Liu et al 1996 [51]

Wargocki 1998 [50]

0%

Sustainable and Healthy Built Environment, Fig. 12 A number of international studies link the quality of building materials and assemblies to allergies, asthma, SBS, and mucosal symptoms

106

composite wood products, as compared to those exposed to formaldehyde from furnishings and products in their home [44]. While sustainable design depends on the use of materials and assemblies that support healthy indoor environments, it also mandates the selection of materials with low-embodied and transportation energy, since these environmental costs carry secondary health concerns. The most rapidly emerging body of research linking interior materials and health may be related to the microbial infections transferred by contact with surfaces, door handles, faucets, keyboards, telephones, even elevator buttons, and the importance of hands-free design and frequent hand washing. One study by Rheinbaben et al. identified that viruses can be transferred to six people successively from contaminated door handles in a household or dormitory setting [50]. Contact infection can be partially addressed through hand washing and green cleaning techniques as well as through hands-free design innovations for shared facilities and equipment. The seriousness of microbial infections from surfaces as well as from water- and airborne sources is the subject of a 2017 US National Academy of Sciences study. Managing acoustic quality in indoor environments is critical to both productivity and human health. The importance of reducing unnecessary noise sources while ensuring appropriate sound transmission must continue to be a major requirement for high-performance and sustainable indoor environments. In a 2002 study of ten volunteer teachers from ten randomly selected preschools in Stockholm, Sweden, Sodersten et al. identify an 11% reduction in vocal strain among teachers in classrooms with background noise levels of 55 dBA, as compared to those in classrooms with background noise levels of 75 dBA [51]. Managing unnecessary background noise and ensuring appropriate and necessary sound distribution are as critical to defining indoor environmental quality (IEQ) as visual, thermal, and air quality. While less tied to today’s definition of sustainability, the importance of ergonomic furniture and space configuration for human health must continue to be emphasized in the design of high-performance indoor environments. Given the growing preponderance of computer-based work today, work surfaces, chairs, keyboards, and mouse design must be

Sustainable and Healthy Built Environment

ergonomically designed to reduce musculoskeletal disorders (MSD). According to a Washington State study, 1.7–3.2% of MSD complaints result in medical costs averaging $22,000 per affected occupant and in many cases permanent consequences for the employee [52]. The CBPD has identified seven international case studies that demonstrate that ergonomic workstations reduce MSD symptoms between 48% and 84% (Fig. 13) [53–59]. Ergonomic design goes beyond anthropometric concerns, however, to also address building layout and densities that support human health and productivity. Jaakkola and Heinonen [60] identified a 35% lower rate of colds among occupants of individual offices, compared to those in shared offices. John Templer [61], the author of a two-volume reference on Stairs, carefully illustrates critical design decisions for stairs, ramps, curbs, and surfaces to reduce the frequency of falls, the most frequent cause of injury and death in buildings. Moving beyond the importance of land use and community design to increased physical activity and health, interior design also plays a major role. The New York City public health department has developed design guidelines to promote physical activity in new and existing buildings in an effort to address obesity and its related diseases (https:// centerforactivedesign.org/impactreport). These Active Design Guidelines have led to a collaboration with the US Center for Disease Control in the development of the FITWEL standard for creating healthier buildings, streets, and urban spaces, based on the latest academic research and best practices in the field. The role of interior architecture, engineering, and design for human health is significant and suggests a rich mix of critical design guidelines for improved human health over time. Linking sustainable material guidelines with health • Specify materials that do not irritate the skin with contact to avoid dermatological conditions • Specify materials that do not outgas toxins to avoid respiratory/allergy and asthma • Specify materials that do not degenerate into respirable fibers or emit radon to avoid cancers • Specify materials that are not fire hazards causing respiratory illness or death • Specify materials that do not foster mold or mildew leading to respiratory symptoms (continued)

Sustainable and Healthy Built Environment

107

Musculoskeletal Disorder Reduction due to Ergonomic Improvements

90%

CBPD ABSIC BIDSTM

100% Ergonomic chair and workstation + ergonomic training 80.0%

82.0%

80.0%

80% 70% 57.0% fixed-split keyboard

63.0% %Reduction

60%

60.0% adjustable keyboard tray

50%

48.0% ergonomic chair + training

40% 30% 20% 10% 0% USGAO 1997 [60]

OSHA 1999 [61]

Ignatius and Fryer 1994 [62]

OSHA [63]

Zecevic and Harburn 2000 [64]

Hedge et al 1999 [65]

Amick et al 2003 [66]

Sustainable and Healthy Built Environment, Fig. 13 A number of international studies link musculoskeletal disorder reduction to ergonomic improvements

Linking sustainable material guidelines with health • Specify materials with low-embodied energy and low transportation costs to reduce outdoor air pollution • Specify furniture ergonomics to reduce musculoskeletal disorders (MSD) • Design spatial layout/density to reduce transmission of contagious illnesses (flus, colds) • Design spatial layout to reduce falls and tripping • Design layout and specify surfaces to reduce infections transferred by contact with hands-free design • Design spatial layout for enhanced physical activity to reduce obesity and related health concerns

Sustainable Maintenance and Operations for Health Needless to say, each of these design guidelines will become obsolete if there is no commitment to longterm maintenance and operational standards. The building enclosure, HVAC, and lighting systems must be continuously commissioned to maintain the healthy conditions intended. Standing water,

dampness, and mold must be prevented. Occupant densities must be managed, and furniture and finishes must continue to meet the latest health standards. In the past few years, important standards, declarations, and certifications relative to environmental and human health of materials and products have emerged for the building design and management community. An EPD® (Environmental Product Declaration) is an independently verified and registered document that communicates transparent and comparable information about the life cycle environmental impact of products. The HPD Open Standard provides specification for the accurate, reliable, and consistent reporting of product contents and associated health information for products used in the built environment. Declare™ is a transparency platform and product database that is changing the materials marketplace by capturing information on where a product comes from, what it is made of, and where it goes at the end of its life. The Cradle to Cradle Certified™ Product Standard guides designers and manufacturers through a continual

108

improvement process that looks at a product through five quality categories — material health, material reutilization, renewable energy and carbon management, water stewardship, and social fairness. In addition, human activities in buildings and the products that occupants bring in must also be selected for health. Cleaning supplies, plants, fertilizers and herbicides, and office and teaching supplies must all be environmentally benign. In addition, food and water quality should be monitored for health, including guidelines for vending machines. Waste must be effectively managed since it is a natural breeding ground for roaches, rodents, and other pests and an opportunity for recycling or composting. While this research team has not evaluated the studies that may link poor maintenance and operation practices to health concerns, it is clear that degradation in as-built performance can result in health consequences equally serious as those of poor design, engineering, and construction.

Sustainable and Healthy Built Environment, Fig. 14 The true cost of business extends well beyond mortgage and energy costs, to include salaries, benefits,

Sustainable and Healthy Built Environment

Calculating the Life Cycle Benefits of Sustainable Design and Health The work of the faculty, researchers, and graduate students of the Center for Building Performance and Diagnostics at Carnegie Mellon and the Advanced Building Systems Integration Consortium extends beyond the pursuit of building case studies that link the quality of buildings to productivity, health, and life cycle sustainability. The development of the BIDS™ tool is based on the identification of economic, environmental, and human life cycle cost benefits related to buildings and communities in order to calculate the return on investment of high-performance building systems (see http://cbpd.arc.cmu.edu/ebids). Figure 14 helps to reveal the diverse building-related costs of doing business in US offices, including salaries and health benefits, technological and spatial churn, rent, energy, and maintenance costs. This cost is normalized in dollars per person per year,

technological and spatial churn. Improving the quality of the built environment can be offset by sustained health benefits

Sustainable and Healthy Built Environment

109

rather than cost per square foot, since the employee represents both the greatest cost and the greatest asset to an organization. The CBPD research identified that across independent nonprofit organizations, human resource research firms, and the US government, the average employer cost for health insurance was approximately $5000 per employee per year in 2003 [62–66]. Within this $5000 expense, the CBPD continues to identify the cost of health conditions and illnesses that can be linked to the quality of the indoor environment, including colds, headaches, respiratory illnesses, musculoskeletal disorders, and back pain (shown in Fig. 14), which account for roughly $750 of the $5000 annually spent per employee or 14% of all annual health insurance expenditures. The direct costs for medical attention and pharmaceuticals would be multiplied further with the indirect costs of reduced speed and accuracy on task and lost work time due to absenteeism. Beyond the organizational and human costs of worker absenteeism, however, Paul Kemp in Harvard Business Review introduced the term “presenteeism” to reflect the more typical and serious condition of individuals coming to work sick because they cannot afford to miss more work hours or important meetings [67]. While the

MEDICAL and PHARMACEUTICAL 24%($116.2M) DIRECT MEDICAL COSTS INDIRECT MEDICAL COSTS PRESENTEEISM 63%($311.8M)

Long-term disability 1% [$6M]

Absenteeism 6% ($27M)

Short-term disability 6% ($27M)

measured costs of presenteeism are not well documented, there is a modest literature to help quantify the effective work time that is lost on the job due to headaches, colds, flus, and back pain (see Fig. 15), [68]. The importance of quantifying the financial, environmental, as well as human health and productivity impacts of design decisions is critical to the advancement of sustainable buildings and communities. In his 1998 book Cannibals with Forks: The Triple Bottom Line of 21st Century Business, John Elkington [69] introduces the importance of adding social and environmental impacts to economic performance in a triple net present value calculation. To accomplish this, it is critical to quantify the environmental costs of using energy and raw materials without caution, as well as the human health and performance costs of poor land use planning and building design and operation. For the past 10 years, the CMU Center for Building Performance and Diagnostics has been building the economic baselines for triple bottom line calculations for offices, schools, and hospitals (see Fig. 16). For example, the calculation of life cycle benefits of better design, engineering, and management of hospitals would include variables such as the average length of stay per illness, at 4.6 days

Condition

Prevalence

Average productivity loss

Aggregate ammual loss

Migraine

12.0%

4.9%

$434,385

Arthritis

19.7

5.9

865,530

Chronic lower-back pain (without leg rain)

21.3

5.5

858,825

Allergies or sinus trouble

59.8

4.1

1,809,945

Asthma

6.8

5.2

259,740

GERD (acid reflux disease)

15.2

5.2

582,660

Dermatitis or other skin condition

16.1

5.2

610,740

Flu in the past two weeks

17.5

4.7

607,005

Depression

13.9

7.6

786,600

Source: Debra Lerner, William M. Rogers, and Mong Chang, & Tofts-New England Medical Center

Source: Back One Figures are based on annual data for 2000 workers’ compensation accounted for less than 7% of indirect medical costs

Sustainable and Healthy Built Environment, Fig. 15 While absenteeism is a cost of business, presenteeism – coming to work with colds, flus, and back pain, for example – may have an even greater impact on the bottom line

110

Sustainable and Healthy Built Environment

Sustainable and Healthy Built Environment, Fig. 16 Measuring “productivity and health” in different building types will require careful identification of critical outcomes and consistent data collection

per patient in US hospitals [70]; average cost of hospital stay, set at $1217 per day in US hospitals [71]; patient reinfection rates, estimated at 2.16/ 10,000 patient days in US hospitals [72]; average cost of these nosocomial infections, estimated at $27,000 plus 12-day increase in hospital stay [73]; and the average cost of nurse turnover, at $13,800 per nurse per year [74–76]. The magnitude of these costs would clearly justify significant investment and reinvestment in the quality of hospitals to ensure long-term health and productivity. Cars and laptops are purchased with far more comprehensive life cycle considerations than buildings, and yet the life span of cars and laptops are often 5 years or less. Since buildings are built for 30, 50, or even hundreds of years, it is imperative that the client and design community begin to address life cycle costs of materials, components, and integrated systems with triple bottom line decision-making to ensure the sustainability of economic, environmental, and human health and productivity. Triple Bottom Line (TBL) accounting can transform the value engineering process by introducing the environmental and human benefits of investing in better building materials and assemblies. For example, to significantly improve visual health and visual performance, the benefits of lowering ambient light levels for today’s dominantly computer-based tasks, and adding task lights for the occasional paper task, result in payback periods shortening from 3 years to 2 years to 8 months (see Fig. 17 [77–79]). The CMU Center for Building Performance and Diagnostics has been developing TBL calculations for a range of

high-performance building systems that impact human health, to encourage investment to move beyond first, least-cost decision-making.

Health and the Built Environment: A Research Mandate Sustainability is in truth all about health. Energy/ material extraction and use as well as atmospheric, water, and land pollution are as significantly healthrelated issues as they are environmental conservation issues. Certainly the design and maintenance of building enclosures, HVAC, lighting, and interior systems are directly linked to our short- and longterm health, as the evidence collected in this chapter has begun to prove. Human health in relation to the built environment is one of the most critically needed research efforts, requiring both extensive experimental and field research efforts. Controlled laboratory experiments need to be carried out simultaneously with experiments in the field – to map chains of consequence – and identify possible building-related causes for respiratory, digestive, circadian, musculoskeletal, circulatory, and nervous system illnesses, as well as other health-related concerns. Yet in the USA, at least, there is remarkably little federal investment in defining and valuing healthy buildings and communities. One cannot overstress the importance of defining key national and international research directions for addressing the impact of the built environment on health. Bringing together emerging knowledge about the importance of land use, building

Sustainable and Healthy Built Environment

111

+ Costs to Reduce Ambient Lighting and Add Task Lights Cost for reducing ambient light levels Cost for LED desk lamp First cost for the investment Initial Investment costs for a 100,000 sq ft building

Per sq ft $0.16 $0.82 $0.98

Per employee $32 $164 $196 $98,000

Per sq ft $0.27 $0.05 +$0.32

Per employee $54 $10 +$64

1st Financial Capital savings Energy savings (40%)1 O & M Savings2 Annual 1st bottom line savings total Cumulative ROI (Financial) Payback Period 15 year Net Present Value

32% 3 years $244,000

2nd Financial + Natural Capital savings Environmental benefits from energy savings of: Air pollution emissions (SOX, NOX, PM)3 CO2 reductions 3 Water savings 4,5,6 Annual 2nd bottom line savings Cumulative ROI (Financial + Natural) Payback Period

Per sq ft 2.71 kWh $0.04 $0.03 $0.01 +$0.08

Per employee 542 kWh $9 $5 $2 +$16 41% 2 years

15 year Net Present Value

$305,860

3rd Financial + Natural + Human capital savings of Reducing Ambient and Adding Task Absenteeism reduction (1%)7 Productivity increase (11%)8 Health benefits (19%)9 Annual 3rd bottom line savings Cumulative ROI (Economic + Environment+ Equity)

Per sq ft $0.03 $0.90 $0.07 $1.00

Per employee $6 $180 $14 $200 140%

Payback Period

8 months

15 year Net Present Value (10% discount rate)

$1,058,550

Sustainable and Healthy Built Environment, Fig. 17 (Triple Bottom Line) Calculations enable decision-makers to see the rapid reductions in payback periods, and the substantial increases in 15 year net present value, given

the health and productivity benefits, in this case for reducing ambient light levels and adding task lights for visual health and performance [77–79]

enclosure, HVAC, lighting, and interior design decisions, with the life cycle justifications to ensure their implementation, is critically needed. Sustainable buildings and communities have the potential to

deliver the highest-quality air, thermal control, light, ergonomics, and acoustic quality, as well as regionally appropriate access to the natural environment, which are integral to human health.

112

In the face of rapid changes in the built environment, the importance of proving that sustainable design and engineering improves health, productivity, and quality of life has never been more important. Indeed, the advances most critically needed for environmental sustainability will constantly be slowed by least first-cost decisionmaking, unless the health-related benefits of sustainable buildings and communities are definitively revealed.

Bibliography 1. Loftness VE (2005) Building Investment Decision Support (BIDS™) – cost-benefit tool to promote high performance components, flexible infrastructures & systems integration for sustainable commercial buildings and productive organizations. AIA pilot report on university research, AIA, Washington, DC, pp 12–31 2. Cornell University Medical College, C. U (1949) A brief history of the cornell medical index. Retrieved from Weill Cornell Medical Library. http:// library.med.cornell.edu/About/cornellmedindex.html. Accessed 27 Dec 2011 3. Catlin T, Simoes E, Brownson R (2003) Environmental and policy factors associated with overweight among adults in Missouri. Am J Health Promot 17(4):249–258 4. Ewing R, Schmid T, Killingsworth R, Zlot A, Raudenbush S (2003) Relationship between urban sprawl and physical activity, obesity, and morbidity. Am J Health Promot 18(1):47–57 5. Giles-Corti B, Macintyre S (2003) Environmental and lifestyle factors associated with overweight and obesity in Perth, Australia. Am J Health Promot 18(1):93–102 6. Saelens BE, Sallis JF, Black J, Chen D (2003) Neighborhood-based differences in physical activity: an environment scale evaluation. Am J Public Health 93(9):1552–1558 7. Frumkin H (2002) Urban sprawl and public health. Public Health Rep 117(3):201–217 8. Skole DL (n.d.) Land use, human health and climate change: convergence of planning problems for Michigan. Retrieved from http://www.landuse.msu.edu/. Accessed 27 Dec 2011 9. Wordley J, Walters S, Ayres J (1997) Short term variations in hospital admissions and mortality and particulate air pollution. Occup Environ Med 54:108–116 10. Dockery D, Pope C (1994) Acute respiratory effects of particulate air pollution. Annu Rev Public Health 15:107–132 11. Tenias JM, Ballester F, Rivera ML (1998) Association between hospital emergency visits for asthma and air pollution in Valencia, Spain. Occup Environ Med 55:541–547

Sustainable and Healthy Built Environment 12. Frank D, Andresen M, Schmid T (2004) Obesity relationships with community design, physical activity, and time spent in cars. Am J Prev Med 27(2):87–96 13. Ewing R, Brownson RC, Berrigan AD (2006) Relationship between urban sprawl and weight of United States youth. Am J Prev Med 31(6):464–474 14. Ewing R, Schmid T (2003) Relationship between urban sprawl and physical activity, obesity, and morbidity. Am J Health Promot 18(1):47–57 15. Beauchemin KM, Hays P (1996) Sunny hospital rooms expedite recovery from severe and refractory depression. J Affect Disord 40:49–51 16. Benedetti F, Colombo C, Barbini B, Campori E, Smeraldi E (2001) Morning sunlight reduces length of hospitalization in bipolar depression. J Affect Disord 62:221–223 17. Choi J (2005) Study of the relationship between indoor daylight environments and patient average length of stay (ALOS) in healthcare facilities. Unpublished master’s thesis, Department of Architecture, Texas A&M University, College Station 18. Walch JM, Rabin BS, Day R, Williams JN, Choi K, Kang JD (2005) The effect of sunlight on postoperative analgesic medication use: a prospective study of patients undergoing spinal surgery. Psychosom Med 67:156–163 19. Bullough JD, Rea M (2006) Of mice and women: light as a circadian stimulus in breast cancer research. Cancer Causes Control 17(4):375–383 20. Ulrich RS (1984) View through a window may influence recovery from surgery. Science 224(4647):420–421 21. Mendell MJ (1991) Risk factors for work-related symptoms in Northern California office workers. Unpublished doctoral dissertation, University of California, CA Dept. of Health Services, and NIOSH, indoor environment program Lawrence Berkeley Lab., CA. Sponsoring Org. USDOE, Washington, DC 22. Heschong Mahone Group (2003) Windows and offices: a study of office worker performance and the indoor environment. Program, funded by the California Energy Commission’s Public Interest Energy Research (PIER) program 23. Kellert SR (2005) Building for life: designing and understanding the human-nature connection. Island Press, Washington, DC 24. Kelland P (1992) Sick building syndrome, working environments and hospital staff. Indoor Built Environ 1:335–340 25. Preziosi P, Czernichow S (2004) Workplace airconditioning and health services attendance among French middle-aged women: a prospective cohort study. Int J Epidemiol 33(5):1120–1123 26. Zuraimi MK (2007) The effect of ventilation strategies of child care centers on indoor air quality and respiratory health of children in Singapore. Indoor Air 17:317–327 27. Bourbeau J, Brisson C, Allaire S (1997) Prevalence of the sick building syndrome symptoms in office workers before and six months and three years after being exposed to a building with an improved ventilation system. Occup Environ Med 54:49–53

Sustainable and Healthy Built Environment 28. Brundage J, McNeil R, Lednar W, Smith D, Miller R (1988) Building-associated risk of febrile acute respiratory diseases in army trainees. JAMA 259(14):2108–2112 29. Drinka P, Krause P, Schilling M, Miller B, Shult P, Gravenstein S (1996) Report of an outbreak: nursing home architecture and influenza – a attack rates. J Am Geriatr Soc 44:910–913 30. Fisk W, Rosenfeld A (1997) Estimates of improved productivity and health from better indoor environments. Indoor Air 7:158–172 31. Jaakkola J, Miettinen P (1995) Ventilation rates in office buildings and sick building syndrome. Occup Environ Health 94(2):709–714 32. Menzies DF, Yuan L, Fitzgerald J (2000) Hospital ventilation and risk for tuberculosis infection in Canadian health care workers. Ann Intern Med 133(10):779–789 33. Myatt T, Johnston S, Zuo Z, Wand M, Kebadze T, Rudnick S et al (2004) Detection of airborne rhinovirus and its relation to outdoor air supply in office environments. Am J Respir Crit Care Med 169:1187 34. Smedje G, Norback D (2000) New ventilation systems at select schools in Sweden – effects on asthma and exposure. Arch Environ Health 35(1):18–25 35. Sundell J (1996) What we know, and don’t know, about sick building syndrome. ASHRAE J 38:51–57 36. Tham KH (2003) Temperature and ventilation effects on the work performance of office workers (study of a call center in the tropics). In: Proceedings of healthy buildings 2003, Singapore 37. Seppänen OA, Fisk WJ, Mendell MJ (1999) Association of ventilation rates and CO2 concentrations with health and other responses in commercial and institutional buildings. Indoor Air 9(4):226–252 38. Sundel J, Levin H, Nazaroff WW, Cain WS, Fisk WJ, Grimsrud DT et al (2011) Ventilation rates and health: multidisciplinary review of the scientific literature. Indoor Air 21(3):191–204 39. Kaczmarczyk J, Zeng Q, Melikov A, Fanger P (2002) The effect of a personalized ventilation system on perceived air quality and SBS symptoms. In: Proceedings of indoor air 2002. Monterey 40. Menzies D, Popa J, Hanley J, Rand T, Milton D (2003) Effect of ultraviolet germicidal lights installed in office ventilation systems on workers’ health and wellbeing: double-blind multiple crossover trial. Lancet 362:1785–1791 41. Menzies D, Pasztor J, Nunes F, Leduc J, Chan C (1997) Effect of a new ventilation system on health and well-being of office workers. Arch Environ Health 52(5):360–367 42. Wilkins AN-S (1989) Fluorescent lighting, headaches and eyestrain. Light Res Technol 21(1):300–307 43. Çakir AE, Çakir G (1998) Light and health: influences of lighting on health and well-being of office and computer workers. Ergonomic, Berlin 44. Garrett M, Hooper M, Hooper B (1996) Low levels of formaldehyde in residential homes and a correlation with asthma and allergy in children. In: Proceedings of the 7th

113

45.

46.

47.

48.

49.

50.

51.

52.

53.

54.

55. 56.

57. 58.

59.

international conference of indoor air quality and climate (Indoor Air996), vol 1. Nagoya, pp 617–622 Jaakkola J, Tuomaala P, Seppanen O (1994) Textile wall materials and sick building syndrome. Arch Environ Health 49(3):175–182 Jaakkola JJ, Oie L, Nafstad P, Botten G, Samuelson SO, Magnus P (1999) Interior surface materials in the home and the development of bronchial obstruction in young children in Oslo, Norway. Am J Public Health 89:188–192 Liu JZ, Tao YX, Hao L (1996) The relationship between sick building syndrome and indoor decoration. In: Proceedings of the 7th international conference on indoor air quality and climate (Indoor Air’96), vol 2. Nagoya, pp 321–324 Wargocki P (1998) Human perception, productivity, and symptoms related to indoor air quality. Doctoral thesis, Center for Indoor Environment and Energy, Technical University of Denmark Wieslander G, Norback D, Bjornsson E, Janson C, Boman G (1997) Asthma and the indoor environment: the significance of emission of formaldehyde and volatile organic compounds from newly painted surfaces. Int Arch Occup Environ Health 79(2):115–124 Rheinbaben F, Schünemann S (2000) Transmission of viruses via contact in a household setting: experiments using bacteriophage fX174 as a model virus. J Hosp Infect 46:61–66 Sordersten MG (2002) Vocal loading factors for preschool teachers at work studied with binaural DAT recordings. Department of Speech Music and Hearing, Royal Institute of Technology, Stockholm Silverstein B, Vilkari-Juntura E, Kalat J (2000) Workrelated musculoskeletal disorders of the neck, back, and upper extremity in Washington state, 1990–1988, technical report number 40-4a-2000. Safety and Health Assessment & Research for Prevention, Washington State Department of Labor and Industries, Washington, DC Amick B, Robertson M, DeRango K, Bazzani L, Moore A, Rooney T et al (2003) Effect of office ergonomics intervention on reducing musculoskeletal symptoms. Spine 28(24):2706–2711 Occupational Safety and Health Administration, U. D (1999) Preliminary economic analysis and initial regulatory flexibility analysis for the occupational safety and health administration’s proposed ergonomics programs standard (Chapter III-V and Appendix III-A, Scenario No. OGI14). US Department of Labor Occupational Safety and Health Administration, U. D (2006) Success with ergonomics Hedge A, Morimoto S, McCrobie D (1999) Effects of keyboard tray geometry on upper body posture and comfort. Ergonomics 42(10):1333–1349 Ignatius EA (1994) The high cost of keyboard injuries. PC World 12(3):45–46 United State General Accounting Office, U. S (1997) Worker protection: private sector ergonomics programs yield positive results. GAO/HEHS-97-163 Zecevic A, Miller DI, Harburn K (2000) An evaluation of the ergonomics of three computer keyboards. Ergonomics 43(1):55–72

114 60. Jaakkola JJ (1995) Shared office space and the risk of the common cold. Eur J Epidemiol 11:213–216 61. Templer J (1995) The staircase: studies of hazards, falls, and safer design. MIT Press, Cambridge 62. Bureau of Labor Statistics, U. D (2002) National compensation survey: occupational wages in the United States 2001. Bureau of Labor Statistics, US Department of Labor, Washington, DC 63. Deloitte and Touche D (2003) Employer health care strategy survey 2003 64. Kaiser Family Foundation and Health Research and Educational Trust K. F (2003) Employer health benefits: 2003 annual survey.Retrieved from Kaiser Family Foundation and Health Research and Educational Trust. http://ehbs.kff.org/. Accessed 27 Dec 2011 65. Towers Perrin HR Services TP (2003) Towers Perrin 2004 health care cost survey 66. US Chamber of Commerce Statistics and Research Center, U. C (2003) 2003 Employee benefits study 67. Hemp P (2004) Presenteeism: at work – but out of it. Harv Bus Rev 82(10):49–58 68. Lerner D, Chang H, Rogers WH, Benson C, Schein J, Allaire S (2009) A method for imputing the impact of health problems on at-work performance and productivity from available health data. J Occup Environ Med 51(5):515–524 69. Elkington J (1998) Cannibals with forks: the triple bottom line of 21st century business. New Society, Gabriola Island 70. Project H (n.d.) Agency for healthcare research and quality. Retrieved from http://hcupnet.ahrq.gov/ HCUPnet.app/. Accessed 27 Dec 2011 71. US Census Bureau, U. C (2003) Statistical abstract of the United States: 2003, 123rdrd edn. U.S. Census Bureau, Washington, DC

Sustainable and Healthy Built Environment 72. Pittet D, Hugonnet S, Harbarth S, Mourouga P, Sauvan V, Touveneau S et al (2000) Effectiveness of a hospital-wide program to improve compliance with hand hygiene. Lancet 356:1307–1312 73. North Carolina Statewide Program for Infection Control and Epidemiology S (n.d.) SPICE (North Carolina statewide program for infection control and epidemiology). Retrieved from http://www.unc.edu/depts/ spice/MRSA-VRE-Surveillance.ppt. Accessed 27 Dec 2011 74. AONE AO (2002) Acute care hospital survey of RN vacancies and turnover rates in 2000. American organization of nurse executives 75. Jones C (2004) The costs of nurse turnover. Part 1: an economic perspective. J Nurs Adm 34(12):362–370 76. Jones C (2005) The costs of nurse turnover. Part 2: application of the nursing turnover cost calculation methodology. J Nurs Adm 35(1):41–49 77. Loftness L, Srivastava R, Cochran E (2013). Triple bottom line benefits of investing in lighting and daylighting. Futurebuild conference proceedings, Bath 78. Loftness V, Srivastava R, Dadia D, Parekh H, Rawal R, (2014). The triple bottom line benefits of climate responsive dynamic facades. PLEA 2014 international conference on Sustainable Habitat for Developing Societies, Ahmedabad 79. Srivastava R (2017). Integrating financial, natural and human capital – the triple bottom line – for high performance investments in the built environment. Doctoral dissertation, Carnegie Mellon University, Pittsburgh 80. CBPD (n.d.) eBIDS – energy building investment building support. Retrieved from center for building performance and diagnostics. School of Architecture; Carnegie Mellon University. http://cbpd.arc.cmu.edu/ ebids/pages/home.aspx. Accessed 27 Dec 2011

Regenerative Development and Design Pamela Mang1 and Bill Reed2 1 Regenesis Group, Santa Fe, NM, USA 2 Regenesis Group, Arlington, MA, USA

Article Outline Definition of Regenerative Design and Its Importance Introduction Regenerative Development and Design: Redefining Sustainability Overview: Ecological Sustainability and Regenerative Development and Design Regenerative Approaches to Sustainable Development and Design: Key Framework Premises and Methods Overview Future Directions Bibliography

Glossary Biomimicry Sometimes called biomimetic design; an emerging design discipline that looks to nature for sustainable design solutions [1]. Cradle-to-cradle Framework for designing manufacturing processes “powered by renewable energy, in which materials flow in safe, regenerative, closed-loop cycles,” and which “identifies three key design principles in the intelligence of natural systems, which can inform human design: Waste Equals Food; Use Current Solar Income; Celebrate Diversity” [2, 3] Ecoliteracy The ability to understand the natural systems that make life on earth possible, including understanding the principles of organization of ecological communities (i.e., ecosystems) and

using those principles for creating sustainable human communities [4, 5]. Ecological sustainability A biocentric school of sustainability thinking that, based on ecology and living systems principles, focuses on “the capacity of ecosystems to maintain their essential functions and processes, and retain their biodiversity in full measure over the long-term”; contrasts with technological sustainability based on technical and engineering approaches to sustainability [4]. Ecology The interdisciplinary scientific study of the living conditions of organisms in interaction with each other and with the surroundings, organic as well as inorganic. Ecosystem concept “A coherent framework for redesigning our landscapes, buildings, cities, and systems of energy, water, food, manufacturing and waste” through “the effective adaptation to and integration with nature’s processes.” It has been used more to shape an approach than as a scientific theory [6, 7]. Ecosystem “The interactive system of living things and their non-living habitat” [6]. Living systems thinking A thinking technology, using systemic frameworks and developmental processes, for consciously improving the capacity to apply systems thinking to the evolution of human or social living systems [8]. Locational patterns The patterns that depict the distinctive character and potential of a place and provide a dynamic mapping for designing human structures and systems that align with the living systems of a place. Pattern literacy Being able to read, understand, and generate (“write”) appropriate patterns. Permaculture A contraction of permanent agriculture or permanent culture, permaculture was developed as a system for designing ecological human habitats and food production systems based on the relationships and processes found in natural ecological communities, and the relationships and adaptations of indigenous peoples to their ecosystems [9].

© Springer Science+Business Media, LLC, part of Springer Nature 2020 V. Loftness (ed.), Sustainable Built Environments, https://doi.org/10.1007/978-1-0716-0684-1_303 Originally published in R. A. Meyers (ed.), Encyclopedia of Sustainability Science and Technology, © Springer Science+Business Media, LLC, part of Springer Nature 2019, https://doi.org/10.1007/978-1-4939-2493-6_303-4

115

116

Place The unique, multilayered network of ecosystems within a geographic region that results from the complex interactions through time of the natural ecology (climate, mineral, and other deposits, soil, vegetation, water, wildlife, etc.) and culture (distinctive customs, expressions of values, economic activities, forms of association, ideas for education, traditions, etc.) [10, 11]. Regenerate American Heritage Dictionary of the English Language and Merriam Webster Dictionary: • To give new life or energy; to revitalize; to bring or come into renewed existence; to impart new and more vigorous life • To form, construct, or create a new, especially in an improved state; to restore to a better, higher or more worthy state; refreshed or renewed • To reform spiritually or morally; to improve moral condition; to invest with a new and higher spiritual nature • To improve a place or system, especially by making it more active or successful Regenerative design A system of technologies and strategies based on an understanding of the inner working of ecosystems that generates designs that regenerate socio-ecological wholes (i.e., generate anew their inherent capacity for vitality, viability, and evolution) rather than deplete their underlying life support systems and resources. Regenerative development A system of developmental technologies and strategies that works to enhance the ability of living beings to coevolve, so that the planet continues to express its potential for diversity, complexity, and creativity [10] through harmonizing human activities with the continuing evolution of life on our planet, even as we continue to develop our potential as humans. Regenerative development provides the framework and builds the local capability required to ensure regenerative design processes achieve maximum systemic leverage and support through time. Regenesis Collaborative Development Group – Restorative design Sometimes called restorative environmental design; a design system that combines returning “polluted, degraded or damaged sites back to a state of acceptable health through human intervention” [12] with

Regenerative Development and Design

biophiliac designs that reconnect people to nature. Source to sink Simple linear flows from resource sources (farms, mines, forests, watershed, oilfields, etc.) to sinks (air, water, land) that deplete global sources and overload/pollute global sinks [13]. Systems thinking A framework for seeing interrelationships rather than things, and for seeing patterns of change rather than static “snapshots.” It addresses phenomena in terms of wholeness rather than in terms of parts [5].

Definition of Regenerative Design and Its Importance The emerging field of regenerative development and design marks a significant evolution in the concept and application of sustainability. Practices in sustainable or green design have focused primarily on minimizing damage to the environment and human health and using resources more efficiently, in effect, slowing down the degradation of earth’s natural systems. Advocates of a regenerative approach to the built environment believe a much more deeply integrated, whole systems approach to the design and construction of buildings and human settlements (and nearly all other human activities). Regenerative approaches seek not only to reverse the degeneration of the earth’s natural systems, but also to design human systems that can coevolve with natural systems – evolve in a way that generates mutual benefits and greater overall expression of life and resilience [10, 11]. The field of regenerative development and design, which draws inspiration from the selfhealing and self-organizing capacities of natural living systems, is increasingly seen as a source for achieving this end. This field is redefining the way that proponents of sustainability are thinking about and designing for the built environment, and even the role of architecture as a field. As an indication of this growing recognition, in May 2017 the SecretaryGeneral of the Commonwealth of Nations (formerly the British Commonwealth), Patricia

Regenerative Development and Design

Scotland, announced the launch of a Commonwealth initiative to reverse climate change through regenerative development, noting that “Regenerative development offers ways of tackling climate change on a scale and by means that can be adopted by the most vulnerable countries, and are appropriate to the day-to-day lives and livelihoods of their inhabitants” [14].

Introduction Early Roots of Regenerative Design In the 1880s, Ebenezer Howard wrote To-morrow: A Peaceful Path to Social Reform. Re-issued in 1902 as Garden Cities of To-Morrow, with an introductory essay by Lewis Mumford, the book was an early and influential expression of ecological thinking applied to human settlement. It sought to reconnect humans to nature, and featured use of natural rather than engineered processes to ensure the health of the system. His description of a utopian city in which man lives harmoniously with the rest of nature stimulated the founding of the garden city movement and the establishment of several Garden Cities in Great Britain in the early twentieth century [13, 15]. In 1915, Patrick Geddes published his study of the urban growth patterns stimulated by the mass movement of people into cities [16]. Geddes, a biologist, saw cities as living organisms. He believed that addressing the problems of unsustainable growth required understanding a city’s context – the surrounding landscape’s natural features, processes, and resources – and called for a solid analytic method for developing that understanding. His conclusion would influence regional planning movements across Europe and the United States. Geddes applied the terms Paleotechnic and Neotechnic to distinguish the industrial era producing this destructive growth of human settlements from the era he predicted would follow its demise. These terms would be picked up by John Tillman Lyle some 80 years later to differentiate industrial era and regenerative technologies. Some trace the origins of ecological design to the work of Patrick Geddes [7, 13].

117

Development of the Ecosystem Concept and Ecological Perspective In 1935, Arthur Tansley introduced an entirely new concept to ecology in his work, “The Use and Abuse of Vegetational Concepts and Terms” [6]. He proposed the term ecosystem as a name for the interactive system of living things and their nonliving habitat, and the application of systems science as a way to bring more scientific rigor to the study of nature’s complexity and the effect of human activities on that complexity. Tansley and other organismic biologists of the period were the first to formulate a systems view of life. Seeking a more accurate depiction of how life ordered and organized itself within a particular landscape or geographic location, he posited that neither a living organism nor its physical environment could be thought of as separate entities: “we cannot separate them from their special environment, with which they form one physical system.” Two of the most significant implications of this depiction of how life structures itself was the deconstruction of the human/nature dichotomy that had shaped Western design thinking, and the establishment of the premise that all species are ecologically integrated with each other, as well as with the abiotic constituents of their biotope or habitat. For Tansley and other ecologists concerned about the increasing impact of humans on natural systems, the ecosystem offered a valuable framework for analyzing the effect of human activities on natural systems and resources. In later years, the concept was further defined or clarified to explicitly include a social complex (human institutions and actions) and a built complex (structures and infrastructures) and became a framework for sustainable urban planning and development [17, 18]. In the 1950s and 1960s, Eugene and Howard Odum laid the foundation for the development of ecology into a modern science, based on the core concept of the ecosystem as the fundamental ordering structure of nature. They published the first textbook on ecology, The Fundamentals of Ecology, in 1953. Their work brought attention to the importance of understanding how the earth’s ecological systems interact with one another. Howard Odum further developed a number of

118

key theoretical concepts and methodologies including his “energy systems language,” a set of symbols used to compose energy flow diagrams for any scale of system. His study of wetlands pioneered the now widespread approach of using wetlands as water quality improvement ecosystems and served as an important contribution to the beginnings of the field of ecological engineering [19]. New Foundations for Systems Thinking In 1968, biologist and systems theoretician Ludwig von Bertalanffy published his General System Theory: Foundations, Development, Applications. General systems theory (GST) introduced the concept of open systems, emphasized the difference between physical and biological systems, and introduced evolutionary thinking – thinking focused on change, growth, and development [20]. GST opened the door to a new science of complexity. The recognition that complex systems cannot be understood through simple analysis led to the emergence of systems thinking as a major scientific field, a profound change from the analytic, reductionist mode that had dominated Western scientific thinking since the time of Descartes, Newton, Galisteo, and Bacon. GST also laid the basis for the development of living systems science. In the 1960s and 1970s, Charles Krone, systems theorist and architect of organizational processes and structures, developed living systems thinking as a developmental technology for consciously improving systems thinking capacity. His work greatly extended GST and Systematics, a discipline developed by mathematician John Bennett that uses systemic frameworks to understand complex wholes within which people are participants rather than observers. The systemic frameworks and developmental processes Krone generated were applied and evolved within businesses. Their purpose was to create an understanding of businesses, communities, and nature as living systems and to build the consciousness required to create reciprocally beneficial relationships through better integration of industrial, community, and natural processes. Krone’s work served as a core foundation for the emerging

Regenerative Development and Design

Regenesis Collaborative Development Group as they developed and evolved regenerative development processes and technologies, starting in the 1990s [10, 21, 22]. Of particular importance in the evolution of regenerative development was Krone’s framework depicting four natures of work that are essential to any living system’s continuing capacity for evolution. The underlying premise is that all four “are necessary in order for an entity to sustain itself in a world that is nested, dynamic, complex, interdependent, and evolving” [10]. The framework defines these different levels of work within a hierarchy (Fig. 1) in which work at the lower levels focuses on existence (what is already manifested), increasing performance, and efficiency. Work at the higher levels is concerned with potential (what could be but is not yet manifested), introducing potential for new life and creativity and advancing the whole. The understanding, aims, and goals developed at the regenerative level work guide work at the other levels. The framework was utilized as an instrument for enabling “practitioners to design for the integrated evolution of all work” and as “a lens for seeing how and where different sustainability strategies fit and how they can be leveraged when aligned around a regenerative goal” [10]. Ecological Sustainability: Foundations of Regenerative Development and Design In 1969, landscape architect Ian McHarg published Design with Nature, pioneering a technology for ecological land-use planning based on

Regenerative Development and Design, Fig. 1 Levels of work. © Regenesis Group (Reprinted with permission)

Regenerative Development and Design

understanding natural systems [23]. His book became a foundational textbook for the ecological view of urban landscape design, and its basic concepts were later developed into today’s geographic information systems (GIS) – a critical tool for ecological development. In 1978, Bill Mollison, an Australian ecologist, and one of his students David Holmgren coined the word permaculture from a contraction of permanent agriculture or permanent culture. They developed the field of permaculture as an ecological design system to promote design of human habitats and food production systems based on the relationships and processes found in natural ecological communities. Much of the inspiration was drawn from the relationships and adaptations of indigenous peoples to their ecosystems. Beyond the integration of human and natural environments, Mollison and Holmgren developed design technologies and practices for increasingly selfsufficient communities and food production systems. By creating “man-made ecosystems,” permaculture demonstrated how to provide for a host of human needs while reducing dependence on environmentally destructive industrial practices. While earlier iterations of ecological design promoted integration of human and natural systems for more sustainable development, permaculture was the first ecological design system to introduce the concept of a regenerative effect as a new standard of ecological performance for the built environment. Peramaculture was based upon the generation of a surplus or overabundance of energy and resources that could be reinvested to evolve natural and human living systems as an integrated whole. In support of that goal, Mollison’s Permaculture: A Designers’ Manual, published in 1988, introduced a hierarchy of investments (regenerative, generative, and degenerative) as a framework for assessing the value of potential actions for building regenerative capacity in a system [9]. Also in the 1980s, Robert Rodale, son of organic agriculture pioneer J. I. Rodale, advanced the use of the word regenerative in relation to the use of land, calling for going beyond sustainability to “where what we are really doing with the American Land is not only producing our food but

119

regenerating, improving, reforming to a higher level the American landscape and the American Spirit” [24]. Rodale used the term to describe the continuing organic renewal of the complex living system that he saw as the basis for healthy soil and, in turn, for healthy food and healthy people. He later applied the same principle of ongoing self-renewal to regenerative economic development [25]. While his work did not extend to the built environment, his principles influenced John Tillman Lyle’s work and are foundational for the conceptualization and application of regenerative methodologies to all of the systems that support life. In 1984, John Tillman Lyle published Design of Human Ecosystems [26] in which he argued that “designers must understand ecological order operating at a variety of scales and link this understanding to human values if we are to create durable, responsible, beneficial designs.” He defined human ecosystems as “places in which human beings and nature might be brought together again” for mutual benefit and posited conscious ecosystemic design as essential to a sustainable future. The book introduced several key concepts that laid the basis for his subsequent work on regenerative design. “Shaping ecosystems, just like shaping buildings” requires (1) a set of organizing principles drawn from “strong concepts of an underlying order that holds the diverse pieces and all their hidden relations together”; (2) “these underlying concepts of order are drawn from ecology,” and principles for ecosystem design “need to comprehend and envision the ecosystem the designer is seeking to shape as a dynamic (living) whole”; and (3) ecological concepts are “more or less analogous to the laws of mechanics in architecture in that they provide us with organizing principles for shaping ecosystems much as architects shape buildings.” Ecological Design Systems Proliferate The 1990s was a period of intense creative ferment for ecological design thinking. A number of foundational books were published laying out both the practical and theoretical bases of design for ecological sustainability, including Ecological literacy: Education and the Transition to a

120

Postmodern World by David Orr (1992), From Eco-Cities to Living Machines: Principles of Ecological Design, by Nancy Jack Todd and John Todd (1993), The Web of Life: A New Scientific Understanding of Living Systems by Fritjoff Capra (1995), Ecological Design by Sim van der Ryn and Stuart Cowan (1996), and The Ecology of Place: Planning for Environment, Economy, and Community by Timothy Beatley (1997). In 1992, Educator David Orr and physicist Fritjof Capra coined the term ecological literacy (also referred to as ecoliteracy) to describe the ability to understand the natural systems that make life on earth possible, including understanding the principles of organization of ecological communities (i.e., ecosystems) and using those principles for creating sustainable human communities [4]. Also in the 1990s, new ecological and living system based metrics were introduced, including architect Malcolm Wells’ Wilderness-Based Checklist for Design and Construction, revised by the Society of Building Science Educators (SBSE) [27]. Their work furthers John Tillman Lyle’s idea that sustainable design might be merely breaking even, while regenerative design renews the earth resources. On a larger scale, Pliny Fisk’s EcoBalance land use planning and design methods employ the principle of life cycles as a framework for sustaining basic life supporting systems, balancing human needs with their ability to enhance the environment, using appropriate technologies for augmenting natural processes [28]. Emergence of Regenerative Development and Design as Distinct Disciplines In 1994, John Tillman Lyle established the Center for Regenerative Design at California State Polytechnic University, Pomona, to test, demonstrate, and further evolve the theory and practice of regenerative design. His book Regenerative Design for Sustainable Development is the first comprehensive articulation of and handbook for regenerative design [13]. Written as a practical guide to the theory and design of regenerative systems, it laid out the framework, principles, and strategies for design aimed at reversing the

Regenerative Development and Design

environmental damage caused by what Lyle called industrial land use practices. The book reflected the continuing evolution of the thinking he had been pursuing as a landscape architect, architect, and educator. Deeply concerned about resource depletion and environmental degradation in “the design of our twentieth century landscape,” Lyle believed that at the core of the growing environmental crises lay the simplification of living systems caused by “paleo” design and technologies (a term he adopted from Patrick Geddes to depict their relative crudity). “Where nature evolved an ever-varying, endlessly complex network of unique places adapted to local conditions,” he wrote, “. . .humans have designed readily manageable uniformity.” This creates relatively simple patterns and forms designed to be easily replicable anywhere. Most important, in his view, was the replacement of nature’s continual cycling and recycling of materials and energy – processes “core to the earth’s operating system” – with one-way linear flows from source to sink. “Eventually a one-way system destroys the landscapes on which it depends,” Lyle observed. “The clock is always running and the flows always approaching the time when they can flow no more. In its very essence, this is a degenerative system, devouring its own sources of sustenance.” The degenerative patterns caused by these linear, one-way flows, he believed, demanded a fundamentally different approach that he named regenerative design. Accordingly, Lyle defined regenerative design as the replacement of linear systems of throughput flows with “cyclical flows at sources, consumption centers, and sinks.” The resulting systems provide for “continuous replacement, through (their) own functional processes, of the energy and materials used in their operation” [13]. Lyle died just 4 years after publication of Regenerative Design for Sustainable Development. While he called redesign of the degenerative systems created by industrial linear flows as the “first order of work,” it is clear from the larger body of his work and other writings [29] that he saw regenerative design as encompassing far more than this basic operational goal, as

Regenerative Development and Design

fundamental as it was. While much attention has been given to his models and techniques for designing self-renewing resource and energy flows, Lyle always saw the heart of his work, and the work of regenerative design, as the conscious design of whole ecosystems. The importance of developing a different nature of thinking as the basis for regenerative design, which was addressed in introductory chapters of the book, was left without further development. The narrow definition of the term regenerative as simply “selfrenewing” came to define the focus of regenerative design for many architects and landscape architects for decades thereafter. In 1995, the Regenesis Group was founded and began developing the theoretical and technological foundation for regenerative development – enabling human communities to coevolve with the natural living systems they inhabit, while continuously regenerating environments and cultures. Regenesis founders had practiced biocentric design, inspired by natural processes, in a variety of arenas for a number of years and knew the power of this approach. They maintain that development projects needed to be sources of ecological health, even “engines of positive or evolutionary change for the systems into which they are built” [30] and that the primary drivers of unsustainable patterns was not being addressed by ecological design systems. They saw environmental problems as symptoms of a fractured relationship between people and the living web of nature and argued that the core issue was cultural and psychological, rather than technological. Like Lyle, they believed that addressing this issue required a fundamental transformation in how humans saw their relationship and role with regard to the planet – moving from the current view of standing apart from and using (or protecting) nature to seeing a “co-evolutionary whole, where humans exist in symbiotic relationship with the living lands they inhabit” [30]. For regenerative design to take hold and be successfully applied, the Regenesis team reasoned, a radical shift in thinking and understanding would be required among design professionals, stakeholders, and all the human inhabitants of a place. They proposed the term

121

regenerative development for the more comprehensive work of creating the conditions and building the capacities required for achieving this shift, with the aim of making development a source of harmonious integration with nature [10, 31, 32]. Regenerative Development and Design Emerge as Significant for the Sustainability Movement While awareness and appreciation of regenerative approaches to sustainability grew through the early 2000s, regenerative development and design continued to be largely an edge phenomenon for much of the decade. In the following decade, however, a series of initiatives in multiple disciplines began to transform practice toward regenerative goals: • Regenerative Built Environment and City Design Studies: In 2010, the World Futures Council published the first of a series of reports introducing the concept of Regenerative Cities, the result of an International Expert Commission on Cities and Climate Change. These reports, followed up by the book “Creating Regenerative Cities” by Herbert Girardet, launched a multi-pronged effort to promote the creation of regenerative cities – cities that work to actively improve and regenerate the productive capacity of the ecosystems on which they depend [33]. • Regenerative Approaches to Economy and Business: JPMorgan Managing Director John Fullerton founded The Capital Institute in 2010, a nonpartisan think-tank dedicated to developing and promoting regenerative economy models. This led to the launch of the Field Guide to a Regenerative Economy and the subsequent white paper, “Regenerative Capitalism: How Universal Principles and Patterns Will Shape the New Economy” as a framework for regenerative economies [34]. In 2016, the Regenerative Business Summit was launched to “elevate and enrich the conversation about regeneration and focus it on innovative enlightened disruption in business and industries” [35]. The following spring, London-based Lush Cosmetics launched the Lush Spring

122

Prize to support the “regeneration movement through an annual 200,000 pound prize fund” and “a high-profile annual conference, bringing people together to share their skills and experiences in raising awareness of regeneration and its potential” [36]. • Regenerative Education Initiatives: In 2013, Regenesis launched “The Regenerative Practitioner,” a blended distance learning series for practitioners interested in integrating regenerative development into their practice. While most course participants were initially professionals within the built environment, the geographic diversity was soon matched by a growing diversity of professions, a reflection of the widening interest in regenerative development. New Zealand hosted the eighth series “for business leaders, cultural leaders, design, development, and planning professionals, researchers, artists, healers, facilitators, community activists, creative entrepreneurs, sustainability managers, permaculturists and community organisers” [37]. Jason McLennan, founder of Living Building Challenge and Living Futures Institute, announced the opening of the School of Regenerative Design in 2017 as “a small, specialized, interdisciplinary design school that focuses on personal transformation, practical knowledgebased internships, grounded in world-class design and ecological thinking” [38]. • Regenerative Design to Address Climate Change: The annual meeting of High Commissioners representing Commonwealth countries adopted regenerative development as the Commonwealth of Nations strategy for reversing climate change. Of the strategy, Patricia Scotland, Secretary-General of the Commonwealth, notes: “Firstly, it is saying that it is possible to reverse the human impact of climate change by 2050 and secondly it is framing climate change as one of our greatest opportunities for innovation and advancement” [39]. Regenerative Design Publications and Conferences: By the mid-2010s, a growing number of journals, books, and conferences, along with dozens of videos on regenerative design and

Regenerative Development and Design

development, reflect the increasing interest in regeneration as a means of reframing sustainable and green practices. Both Building Research and Information journal and the Journal of Clean Production published special issues on the theory andpractice of regenerative design and development and the regenerative sustainability paradigm [40, 41]. Book titles from this period include: 2010 2013

2014 2015 2015 2016 2016 2017

Urban Regeneration and Social Sustainability: Best Practice from European Cities [42] Regenerative Sustainable Development of Universities and Cities: The Role of Living Laboratories [43] Creating Regenerative Cities [33] The Permaculture City: Regenerative Design for Urban, Suburban, and Town Resilience [44] Designing for Hope: Pathways to Regenerative Sustainability [11] Designing for Regenerative Cultures [45] Regenerative Development and Design: a Framework for Evolving Sustainability [10] The Regenerative Business: Redesign Work, Cultivate Human Potential, Achieve Extraordinary Outcomes [46]

Regenerative Development and Design: Redefining Sustainability Introduction Sustainable development and design has been described as falling broadly into two streams – one primarily technical and engineering based (technological sustainability) and the other based in ecology and living systems principles (ecological sustainability) [4, 7]. Green or highperformance building, sometimes called ecoefficient design, emerged out of the first stream, and regenerative development and design out of the second. Green building, like the conventional building field before it, defined the built environment as “all the structures people have built when considered as separate from the natural environment” (MacMillan Dictionary). Green movements defined a sustainable built environment as one that is resource efficient and has minimal or neutral environmental impact. While that definition is evolving, the primary aim of green building continues to be increasing the efficiency of energy, water, and

Regenerative Development and Design

123

material use while reducing local and global impacts on the natural environment. In the past decade, however, the definition of a sustainable built environment is changing rapidly. Sarah Jenkin and Maibritt Pedersen Zari proposed in “Rethinking the Built Environment” that “While aiming for neutral or reduced environmental impacts in terms of energy, carbon, waste or water are worthwhile targets, it is becoming clear that the built environment must go beyond this. It must have net positive environmental benefits for the living world” [12]. The rising field of regenerative development and design, which emerged from the ecological stream, is not only leading the charge to redefine sustainability, but also to redefine what the built environment encompasses and what its role must be. Advocates of a regenerative approach to the built environment believe that a much more

comprehensive, deeply integrated, and wholesystems approach is needed. They propose that eco-efficient design technologies and strategies must be integrated within an ecologically based approach that reverses the degeneration of both the earth’s natural systems and the human systems that inhabit them. The regenerative methodology focuses on the development of human settlements that partner with natural systems and processes to actively regenerate the health of their place as a whole and the spirit of the people who inhabit it (Fig. 2). The philosophical and technical foundations for regenerative development and design as a distinctive field within ecological sustainability were laid in the 1990s, though they draw from scientific and technological advances reaching back into the early part of the last century previously outlined. Held together by a common philosophical core,

An Ecology Habitat-People-Buildings-Infrastructure A Whole Living System

Qualitative Pattern thinking Living & Whole systems Effective-doing the right things Living System Design

ALL IMPROVEMENTS AT ANY OF THESE STAGES ARE IMPORTANT.

Less Energy required. Less initial cost; Less operating cost.

RESTORATION AND REGENERATION CANOT BE ACCOMPLISHED WITHOUT REDUCING THE DAMAGE; YET REDUCING THE DAMAGE IS AN INADEQUATE RESPONSE ON ITS OWN.

Conventional

Green

Buildings Degenerating Infrastructure

Sustainable

Restorative

Regenerating

Living Systems/ Consciousness

Regenerative

More Energy required.

Technical System Design

Efficiency - doing things right Technologies & techniques Fragmented thinking Quantitative

Greater Cost

Trajectory of Ecological Design © All rights reserved. Regenesis 2000-2017 - Contact [email protected] for permission to use

Regenerative Development and Design, Fig. 2 Trajectory of ecological design. © Regenesis Group (Reprinted with permission)

124

regenerative practices extend beyond the traditional aspects of design to address a different nature of thinking and interactivity that is required to design and engage in a regenerative process. While regenerative approaches are attracting growing interest among sustainability design practitioners, transitioning from green building to a regenerative practice has presented a number of challenges. The holistic and deeply integrated nature of the regenerative approach does not lend itself to a “menu approach” – selecting several regenerative technologies without understanding the underlying principles that assure a regenerative outcome. Another challenge is reconciling the two radically different worldviews shaping technological and ecological sustainability within the way one’s practice is carried out. Few architects and engineers are familiar with, let alone trained in an ecological paradigm. Yet as David Orr notes: Ecological problems are in many ways design problems: our cities, cars, houses, and technologies often do not fit in the biosphere. Ecological design requires the ability to comprehend patterns that connect, which means looking beyond the boxes we call disciplines to see things in their larger context. Ecological design is the careful meshing of human purposes with the larger patterns and flows of the natural world; it is the careful study of those patterns and flows to inform human purposes. Competence in ecological design requires spreading ecological intelligence—knowledge about how nature works. [47]

Overview: Ecological Sustainability and Regenerative Development and Design Ecological Sustainability Ecological sustainability has been defined as the “capacity of ecosystems to maintain their essential functions and processes, and retain their biodiversity in full measure over the long-term” (www.businessdictionary.com). While accurate and straightforward, the seeming simplicity of this definition is deceptive. To understand and then deliver what is required to “maintain” and “retain” requires first understanding the nature of ecosystems and the nature of the ecological world in which they exist. That, in turn, requires understanding the ecological perspective – the

Regenerative Development and Design

use of ecological concepts from biology as a metaphor for understanding and designing environments. All development of the built environment involves changing the landscape and, perforce, the natural systems embedded within it – modifying and adapting them for human purposes. The design of that change is ultimately based on the designer’s understanding of the “nature of nature” – how nature works and, concomitantly, humans’ relationship to it. That understanding, in turn, is shaped by the fundamental model or paradigm held by the larger culture and how it understands nature [7, 13, 26, 40, 41]. Divergent ecological versus technological approaches to sustainability can be attributed in large part to their being grounded in very different worldviews. Ecological sustainability, and the design systems within it, emerged from the profound shift in worldview that occurred over the last century as a result of advances in both the physical and biological sciences. Fritjof Capra has described this as a shift from the mechanistic worldview of Descartes and Newton. In the mechanistic paradigm, the dominant metaphor for understanding the world (and all organisms within it) was that of a machine composed of separate parts. In contrast, the ecological worldview sees the world as a self-organizing, continuously evolving, interdependent web of living systems, and the concept of ecosystem is the dominant metaphor for understanding the world. The ecosystem concept, as it has been evolved by living systems science, has been particularly influential in shaping an ecological and regenerative understanding of the world and the role of humans within it, with profound implications for sustainability and development [5, 17, 18]. The industrial-era metaphor of machine was particularly influential in shaping much of the built environment in the developed world and continues to play a significant role even today. By the first decade of the twenty-first century, however, Le Corbusier’s image of the modern house as a “machine for living” was being challenged by the image of living buildings and communities as ecosystems.

Regenerative Development and Design

As the ecosystem emerged as a new “governing concept of relationship between humanity and nature” [17], it confronted some of the most basic premises of the technologies, processes, and goals of the design field at the time, including the role of buildings, the definition of the built environment, the role of designers, and even the role of humans on the planet. As designers concerned about sustainability began to explore the implications of this new paradigm, it became clear that new ways of thinking and working, along with new forms of design and development and new standards of ecological performance were required. Some of the most comprehensive articulations of the key premises that shape the distinctive character of the field of ecological sustainability can be found in the writings

125

of Sim Van der Ryn, Stuart Cowan, David Orr, and Fritjoff Capra [4, 5, 7] Regenerative Development and Design Ecological strategies for sustainability developed during the 1980s and 1990s were organized around the core set of philosophical, theoretical, and scientific concepts. All were aligned around a commitment to net positive goals for the built environment, and to that end were committed to integrating human structures, processes, and infrastructures with natural living systems. To some extent, they differed in the systemic scope they encompassed, falling into four broad categories along a spectrum of comprehensiveness (Fig. 3).

Regenerative Development and Design, Fig. 3 Levels of ecological design. © Regenesis Group (Reprinted with permission)

126

1. Biophilic – As a design philosophy addressing the “urge to affiliate with other forms of life” [48], biophilia is relational in its approach – it is somewhat passive in its engagement with life and is anthropocentric in its purpose. It acknowledges that humans will, if given a choice between nature and a human-made context, choose an environment or situation that utilizes, or is in contact with, living systems and their processes. Human health is positively influenced by connectivity and diminished if separated from living system connectivity. The design fields that employ biophilic approaches consciously use: Physical Engagement and Connections to natural features and elements; Facsimile Connections in terms of the use of nature imagery and materials; and Evocative Connections that use the qualities and attributes of nature in design such as sensory variability, prospect and refuge, serendipity, discovered complexity [49]. 2. Biomemetic – Cradle to cradle and biomimicry are design philosophies that look to nature as inspiration. They are a functional approach that uses nature – its forms and its processes – as a model for humans to follow, an anthropocentric perspective. Technical product design, buildings, manufacturing processes, agriculture, and human activity will function best and be more in harmony with ecological processes if nature is used as a model and guide. Nature’s services and techniques are generally much more effective and certainly more sustainable than technical engineering approximations [50]. The principles guiding biomemetic thinking are essentially derived from an ecological understanding of how life works and provide a conceptual starting point to move into more comprehensive and regenerative systems. 3. Restorative – Restorative approaches seek to improve current systemic performance, returning living systems to a state of health, and reestablishing the self-organizing capability required to maintain that health. This is an approach that acknowledges that humans have a role to play. It is more highly integrated than biomemetic approaches and more active than

Regenerative Development and Design

biophilic approaches – yet it generally is an episodic and finite engagement. This approach typically intervenes on an initial basis to reestablish the health of a subsystem of an ecosystem and community – such as wetlands, woods, riparian corridors, beach dune systems, social systems, and so on. It is a biocentric approach. When the intervening human role is finished however – once the capacity of the system to self-organize is set in motion – the humans leave the engagement [51]. 4. Regenerative – Regenerative approaches embed the capacity to continue to improve performance through time and through varying environmental conditions. Regenerative development and design, as articulated by Regenesis Group and Lyle, recognizes that “humans, human developments, social structures and cultural concerns are an inherent part of ecosystems,” making humans integral, and particularly influential participants in the health and destiny of the earth’s web of living systems. According to this view, the sustainability of the real estate development industry, which works directly on these webs, is largely determined by whether humans participate in them as partners or as exploiters [10]. This might be termed a process of biobecoming – the development of a whole system of inter-related living consciousness – a new mind. “Design inevitably instructs us about our relationships to nature and people, that makes us more or less mindful and more or less ecologically competent. The ultimate object of design is not artifacts, buildings, or landscapes, but human minds” [4]. M. Kat Anderson supports this way of being in “Tending the Wild”: Wilderness is a negative label for land that has not been taken care of by humans for a long time . . . California Indians believe that when humans are gone from an area long enough, they lose the practical knowledge about correct interaction, and the plants and animals retreat spiritually from the earth or hide from humans. When intimate interaction ceases, the continuity of knowledge passed down through generations is broken, and the land becomes “wilderness.” [52]

Regenerative Development and Design

Together, regenerative development and design provide a framework for creating, applying, adapting, and integrating a blend of modern and ancient technologies to the design, management, and continuing evolution of sustainable built environments, accomplishing positive ecological and social results that include: • Improving the health and vitality of human and natural communities – physical, psychological, economic, and ecological • Producing and reinvesting surplus resources and energy to build the capacity of the underlying relationships and support systems of a place needed for resilience and continuing evolution of those communities • Creating a field of caring, commitment, and deep connection to place that enables the changes required for the above to take place and to endure and evolve through time [10, 30]. The first comprehensive articulation of the theoretical and practical basis of regenerative approaches to the built environment emerged separately for regenerative development and regenerative design in the mid-1990s, from two separate sources – the work of Regenesis Group and John Tillman Lyle. Their respective bodies of work each reflected a convergence of disciplines in addition to architecture, including: landscape ecology, geohydrology, landscape architecture, permaculture, regenerative agriculture, general systems theory and cybernetics, living systems theory and thinking, and developmental psychology. In his paper, “New Context, New Responsibilities: Building Capability” [53], Ray Cole articulated some of the key implications of a regenerative approach, including: • Seeing the responsibility of design as “designing the ‘capability’ of the constructed world to support the positive co-evolution of human and natural systems” versus designing “things” (buildings, infrastructure, etc.), and defining sustainable buildings as “buildings that can support sustainable patterns of living.”

127

• Emphasizing the “role of building in positively supporting human and natural processes” versus “building as product.” • Positioning “building as central in creating higher levels of order and, as such, creating increased variety and complexity.” • Seeing the building as within and connected to a larger system – place, shifts “the current emphasis of greater energy self-reliance at the individual building level” to “opportunities for positive connections and creative synergies with adjacent buildings and surrounding natural systems.” A Distinction Between Regenerative Development and Regenerative Design For ecological sustainability to succeed, it requires a far broader and deeper scope of engagement than an individual building or even community design [54]. Yet the structure of the development and construction industry, for the most part, works to narrow the designers’ role and scope, often as a result of decisions made before the design process even begins. Regenerative development was developed as a discipline in part to address this concern. Regenerative approaches view development and design as two distinct yet synergistic processes, both of which play an essential role in ensuring that greater scope, neither of which is sufficient without the other. The following dictionary definitions provide insight into the different roles of development and design: Development: O.Fr. desveloper, “an unfolding, bringing out the latent possibilities,” from des- “undo” + veloper “wrap up” a state in which things are improving; the act of improving by expanding or enlarging or refining; progression from a simpler or lower to a more advanced, mature, or complex form or stage; an unfolding; the discovering of something secret or withheld from the knowledge of others; disclosure. Design: L. designare “mark out, devise,” from de“out” + signare “to mark,” an act of working out the form of something; to create or contrive for a particular purpose or effect.

128

“Regenerative development provides an integrated conceptual framework through which human communities can grow their shared understanding of the unique places in which they live and work. This understanding provides the armature for creating a system of sustainable design strategies and processes tailored to the unique character of a place” [10]. Jenkin and Zari, in their study, “Rethinking the Built Environment,” write that “Regenerative development. . .investigates how humans can participate in ecosystems through development, to create optimum health for both human communities (physically, psychologically, socially, culturally and economically) and other living organisms and systems” [12]. They describe regenerative development as defining the desired outcome in terms of new systemic capabilities, and regenerative design as the means of achieving it. In contrast, John Tillman Lyle [26] defined design within the context of the built environment as giving form to physical processes, and regenerative design as the replacement of linear systems of throughput flows with “cyclical flows at sources, consumption centers, and sinks.” The resulting systems provide for “continuous replacement, through (their) own functional processes, of the energy and materials used in their operation.” Regenerative development works at the intersection of understanding and intention, generating the patterned, whole-system understanding of a place, and developing the strategic, systemic thinking capacities and the stakeholder engagement that are required to ensure designs and design processes achieve maximum systemic regenerative leverage and support. To that end, it integrates building, human and natural development processes within the context of place. Regenerative development also creates an environment that greatly enhances the effect and effectiveness of restorative and biomimetic designs. The roles of regenerative development, more specifically, are to: 1. Develop the whole-systems understanding of the inner working of ecosystems in a specific place required to determine the right phenomena to work on or to give form to, in order to

Regenerative Development and Design

inform and provide direction for regenerative design solutions that can realize the greatest systemic potential; and 2. Build a field of commitment and caring in which stakeholders step forward as cocreators and ongoing stewards of those solutions. Regenerative design solutions that are grown from the uniqueness of a place rather than from a set of universal best practices regenerate rather than deplete underlying life support systems and resources, and work to integrate the flows and structures of the built and natural world “across multiple levels of scale, reflecting the influence of larger scales on smaller scales and smaller on larger” [55, 56].

Regenerative Approaches to Sustainable Development and Design: Key Framework Premises and Methods Overview Key Premises The following four premises are drawn from the work of Regenesis and Lyle. They offer key elements for framing regenerative approaches [10, 13, 21, 26, 31, 32, 57]. The four premises work as a system to integrate and align motivation and means, providing the framework within which methodologies and approaches from other ecological design systems can be integrated into a regenerative practice (Fig. 4). The first two define and shape motive and motivation in a regenerative project. The last two relate to how a project is carried out to ensure that ends and means stay congruent, that the process stays on course toward a regenerative result. 1. Place and Potential – Regenerative projects are based on the richest possible understanding of the evolutionary dynamics of a place in order to identify the potential for realizing greater health and viability as a result of human presence in that place [58]. 2. Goals Focus on Regenerative Capacity – Regenerative projects are defined by the capacity that must be developed and locally

Regenerative Development and Design

Direction: Progressive Harmonization

Ground:Place and Potential

Motivation

Means

Regenerative Development and Design, Fig. 4 Framework depicting key premises and processes characterizing regenerative approaches. © Regenesis Group. Reprinted with permission

129

Goal: Regenerative Capacity

Instrument: Partnering Place embedded to support ongoing coevolution of the built, cultural and natural environments, and the humans who utilize and tend to them – toward higher (more complex, diverse, and generative) levels of order for all their constituent members, as well as for the larger systems they are a part of and dependent on [10, 57]. 3. Partnering with Place – Regenerative projects require taking on a new role, moving from a “builder of systems we control” to a gardener, working in partnership with a place and its processes [10, 21]. 4. Progressive Harmonization – Regenerative projects seek to catalyze a process of continually increasing the pattern harmony between human and natural systems across scales and require indicators and metrics that can track dynamic, holistic, and evolving processes [59]. Place and Potential Potential: ‘the inherent capacity for growth, development or coming into being.’ (American Heritage Dictionary of the English Language)

William McDonough often describes design as an expression of human intention. Both that intention and the resultant design, however, are shaped by the potential the designer sees and seeks to realize for a particular project. Regenerative potential is defined as the ability to leverage human interventions to achieve greater systemic health through time – for the place they occupy and depend on [31]. Many projects fail to achieve a regenerative effect because the potential they target is too limited – focused on an element or a problem without seeing its systemic connections. Others fail because they seek to realize a potential defined

by human ideals but fail to align with the essence of a place and the larger patterns of life that make that place work. When a project is grounded in a rich patterned understanding of its place, and a vision of its role and potential within that place guides its design, even small interventions can ripple out into large systemic transformations – what Curitiba’s long-time mayor Jaime Lerner called “urban acupuncture” [60], with ecological as well as social and economic ramifications. “Place” in regenerative development is alive, a living system or entity that is “. . .a unique constellation of patterns nested within patterns, interwoven with other patterns in families and guilds and social relationships, all endlessly changing, cycling, evolving and building to greater levels of complexity over time. . .an incredibly dynamic and complex being” [59]. A unique, multilayered dynamic network of natural and human ecosystems within a geographic region, this network forms a socioecological whole that is the result of complex interactions through time between and within its constituent ecosystems. The natural ecosystems include wildlife and vegetation, local climate, mineral and other deposits, soil, water, geologic structures, etc.; human ecosystems include distinctive customs, expressions of values, economic activities, forms of association, ideas for education, traditions, physical artifacts such as buildings and constructed infrastructure, etc. [10, 13, 30, 57, 61, 62].

Regenerative Capacity: Defining Goals for Realizing Regenerative Potential The central element for regenerative development and design is the performance not of a single

130

building, but rather of its living context – the unique socioecological system or “place” in which the building is just one of many interdependent and interactive elements and dynamics. Within that context, regenerative goals are set and performance measured in terms of the intended contribution of the built environment to the regenerative capacity of that larger living context – (i.e., its capacity to realize and express more of its full potential as a source of increasingly healthy life for all its constituent members as well as for the larger systems it is a part of and depends on). Characteristics of a regenerative goal include: • Place sourced and place specific • Evolutionary, going beyond improving current systemic performance (what is often called restorative) to embedding into the system the capacity to continue to improve performance through time and through varying environmental conditions • Beyond functional performance goals. Recognizing “human aspiration and will as the ultimate sustaining source of our activities” [27], regenerative goals address qualitative and spirit dimensions that shape the quality and degree of caring humans bring to their place and its capacity to continue to thrive • Focusing on the processes physical structures enable as central Growing Capacity Versus Producing Things: Regenerative projects set place and project specific goals that address all three aspects of regenerative built environments: • Operational capacity • Organizational capacity • Aspirational capacity Operational Capacity Goals: Operational goals focus on systemic functional effectiveness in growing the potential of the underlying resource base – energy, materials, and support systems that enable the evolution of life in a place. Regenerative projects set goals for ensuring that the energies and nutrients flowing through it

Regenerative Development and Design

are used and invested optimally to grow the health of the system and all the life it supports. Organizational Capacity Goals: Organizational capacity focuses on “who” a place is, and addresses two dimensions – what is core to how this place works as a living system (what we can “mess” with and what we can not), and what is the core qualitative character (its essence or distinctiveness) or nature that humans can connect to at a heart level. Goals for this aspect deal with how to utilize the built environment and the design process to both illuminate and enhance the distinctive character of a place as something to be cherished. Historic codes and zones are often used to this end, but they tend to focus on surface appearance rather than essence, and over time the code and its restrictions come to take center stage, overshadowing the living core of the place they intended to protect [21, 63]. Aspirational Goals: Growing the systemic regenerative capacity of a place requires an integration of human aspirations with the distinctive ecosystems of that place and their drive to evolve their own health and generativity. This means harnessing inherent human creativity and aligning it with the creativity of nature, and creating opportunities for people to experience themselves as able to make significant and meaningful contributions to their place [13]. Partnering with Place: a New Role for Humans and Buildings In an ecological paradigm, sustainability requires a fundamental shift in how humans conceive of and carry out their role on the planet. In the words of Joshua Ramo, people must “change the role we imagine for ourselves from architects of a system we can control and manage to gardeners in a living, shifting ecosystem. For hundreds of years now we have lived in our minds as builders: constructing everything from nations to bridges. . . In a revolutionary age, with rapid change all around us, our architects’ tools are deadly. It is time for us to put them down and follow (Nobel Laureate Friedrich von) Hayek’s injunction to live and to think as gardeners.” – gardeners who see themselves as partners in coevolution with the living system in which they

Regenerative Development and Design

work, cultivating “growth by providing the appropriate environment, in the manner a gardener does for his plants” [64, 65]. Successful regenerative development ultimately requires all the stakeholders in a place, not just the development/design team to move from the role of “builder” to “partner–gardener.” To this end, a premise of regenerative development is that “Projects should be vehicles for catalyzing the cooperative enterprises required to enable evolution.” These cooperative enterprises work to bring together stakeholders – people and groups with a stake in growing the potential of their places, around what Stuart Kauffman called “coevolving mutualism,” a progressive and mutually beneficial harmonization of human and natural systems [10, 56]. Partnering with place also requires understanding place as a living whole. Regenerative development starts with a whole systems assessment that looks at a wide range of patterns covering multiple scales of systems and a number of different facets. The place intelligence it develops is a resource that can be mined to inform each stage of design to help ensure that the patterns generated by the project harmonize with the larger patterns of place. To generate the experience of connection and caring that creates a partnership, an understanding of “who” a place is as a living being – its distinct spirit and ways of working – is needed in addition to how it functions. Every living system – whether a person, a tree, or a place – has an ongoing and distinctive core from which it organizes the complex arrays of relationships that produce its activities, its growth, its evolution. Being able to grasp and share the distinctive core or essence of a place among and between the design team and local stakeholders provides an enduring basis for strong partnering relationships, in the same way it builds strong human partnerships. • Regeneration Is a New Way of Thinking: Learning how to apply a regenerative approach begins not with a change of techniques but rather with a change of mind—a new way of thinking about how we plan, design, construct, and operate our built environment. [31]

Growing stakeholders and designing and constructing projects that can work as “place

131

gardeners” requires bringing and developing whole systems thinking that is capable of comprehending, ordering, and organizing the systemic complexity and dynamism of a living place and its multiple scales of nested systems, interactions of multidisciplinary teams over extended periods, and extensive local stakeholder participation [17, 18, 21]. This nature of systems thinking is characterized by: – Being grounded in ecoliteracy and pattern literacy. Ecoliteracy applies an understanding of the fundamental principles that govern how living systems work to specific situations and conditions. Pattern literacy involves being able to read, understand, and generate appropriate patterns that harmonize with and enable a place and its inhabitants to more fully realize what they can be [59]. – Requiring the practitioner to see what they are working on as a system of energies or life processes, rather than as things – to illuminate the reach toward being more whole and alive, a state inherent in living systems that is the fuel for regeneration [32, 66]. – Enabling a diversity of participants to grow their own systems thinking capacity in order to take on more challenging, value-adding roles [10, 22, 66]. • Regeneration Is a New Way of Working: Regenerative development and design does not end with the delivery of the final drawings and approvals, or even with build out of a project. The responsibility of a regenerative designer includes putting in place during the development and design process, what is required to ensure the ongoing regenerative capacity of the project, and the people who inhabit and manage it. Regenerative development employs developmental design processes that encompass integrative design (integrative and interdisciplinary beyond traditional building disciplines, open and participatory), and go beyond to embed self-managed learning processes into the work of conceptualizing, designing, constructing, managing, and evolving regenerative projects [10]. These design processes integrate the traditional organizing for task accomplishment with the development

132

Regenerative Development and Design

of new thinking capacities required to design processes not things, make ecologically sound place-appropriate decisions. They create the connection to and emotional resonance with place that generates the will to follow through on regenerative development and design decisions.

continuous re-patterning. Theoretical biologist Stuart Kauffman called this mutually beneficial relationship “co-evolving mutualism” – coevolving because its ecosystems are always in the process of self-organization and reorganization, increasing in complexity, definition and information content” [26, 69, 70].

Progressive Harmonization The “pole star” or overarching source of direction for regenerative projects derives from the ultimate effect every regenerative project seeks to achieve: an enduring and mutually beneficial relationship between the human and natural systems in a particular place. Pattern is the language of relationship and regenerative development and design in a living system is a process of patterning human communities to align with the energetic patterns of a place in a way that both humans and the place coevolve. Christopher Alexander was speaking of pattern harmony when he wrote “When you build a thing, you cannot merely build that thing in isolation, but must also repair the world around it, and within it, so that the large world at that one place becomes more coherent, and more whole; and the thing which you make takes its place in the web of nature, as you make it” [67]. While his initial work focused primarily on the pattern relationship between a building and the human community and life surrounding, his later work has increasingly encompassed all living systems. Wendell Berry, in his essay “Solving for Pattern,” speaks to creating pattern harmony between human communities and activities and the biosphere they take place in [68]. “A bad (design) solution is bad,” Wendell Berry notes, “because it acts destructively upon the larger patterns in which it is contained. . .most likely, because it is formed in ignorance or disregard of them. A good solution is good because it is in harmony with those larger patterns. . . A bad solution acts within the larger pattern the way a disease or addiction acts within the body. A good solution acts within the larger pattern the way a healthy organ acts within the body” [68]. Pattern harmony however is not a stable state; a good solution today may become a bad one in a few years, so solving for pattern requires a progressive rather than one-time harmonization, a

Regenerative Practice Methodologies Regenesis, collaborative members explored, practiced and evolved a regenerative development methodology over 20 years of fieldwork. The diagram in Fig. 5 was developed as a depiction of the essential phases and developmental processes that are considered key to a regenerative practice that creates and sustains an evolutionary spiral, growing systemic capacity as it actualizes a project [21]. The Three Key Phases of Regenerative Practice: • Understand the Relationship to Place: Integral Assessment – a whole systems (cultural, economic, geographic, climatic, and ecological) assessment of site and place as living systems lay the foundational understanding and thinking required to see how humans can enable the health and continuing evolution of the place and themselves as a part of it. A Story of Place ® is codeveloped with the client and/or community. It uses the power of story telling to articulate the essence of a place, how it fits in the world, and what the role of those who inhabit it can be as collaborators in its evolution (reference). • Design for Harmony with Place: Translate this understanding into design principles and systemic, integrated plans, designs, and construction processes that optimize the presence of people in a landscape by harmonizing human activities with the larger pattern of place. Buildings and infrastructure improve land and ecosystems, and the unique attributes of the land improve the built environment and those who inhabit it. Synergy with the land and ecosystems leverages the effectiveness of green design features and technologies and lowers costs while improving ecosystem health and productivity.

Regenerative Development and Design lution

co-evo

SUSTAIN

design for harmony

HARMONIZE bu

ild

the

fo un

dation

s sse roce p g in anag tive m integra whole systems thinking

UNDERSTAND

growi ng st akeh olde rs

Regenerative Development and Design, Fig. 5 Regenerative practice methodology framework. © Regenesis Group. Reprinted with permission

133

© Regenesis Group

• Design for Co-Evolution: In the words of the Urban Sustainability Learning Group “. . .sustainability means maintaining the dynamic potential for further evolution. Living systems survive by maintaining a condition of dynamic equilibrium with the environment through constant change and adaptation. In the game of evolution, equilibrium is death” [71]. This phase unfolds during and from the work of the previous two phases. If they have succeeded in creating a culture of coevolution in and around the project, and not just a physical product, its effect can be seen even before final build out. The role of designer becomes one of resource, providing processes and methods for sustaining the connection to place as a context that enables owners, managers, contractors, and community stakeholders to recognize and incorporate new social, economic, and ecological opportunities as their place evolves. The Three Key Development Processes in Regenerative Practices: Success in the above three steps is determined by how we think, how we identify harmonies and harmonize the human role, and how we engage

stakeholders throughout the planning and development process. Specifically, through: • Applying whole-systems thinking to the design, planning, and decision-making processes • Managing integration and harmonization across disciplines, between phases and team members and local stakeholders • Growing stakeholders understanding and appreciation of the place and the new potential offered, and their capacity to be increasingly effective partners with the system of evolving life Illustrations of Regenerative Practices The three key phases of regenerative practice – understanding the relationship to place, designing for harmony with place, and designing for coevolution – capture the richness of the precedents in regenerative thinking described in this entry. The following paragraphs illustrate regenerative thinking and practice frameworks and methodologies applied within the three phases, some developed by Regenesis, some drawn from other ecological design systems. In understanding the relationship to place, the principles from permaculture and biomimicry are

134

Regenerative Development and Design

helpful in developing specific land use, building, and infrastructure design strategies. As a design system rooted in the ability to discern the patterns that are structuring both natural and human systems, and to generate new patterns that weave the human and nature together into a dynamic whole, permaculture assessment methodologies provide a source for developing holistic site assessments. “Pattern as Process,” an article by Regenesis principal Tim Murphy and Vickie Marvick, provides a detailed description of their method for understanding and interpreting the patterning of a site and its place [59]. The challenge is to ensure that the scope being assessed is whole enough to encompass the interweaving of human and natural systems, dynamics and flows that shape the distinct character of a place (Fig. 6). Regenesis developed the following framework as a means of illuminating the core patterns structuring a place as the basis for “mapping” their dynamic and evolving interrelationships. These patterns include:

• The ecological, social, and cultural systems creating and managing the conditions that shape how life expresses itself in a place • The value adding processes that life engages in within the context of those conditions and how they influence and are influenced by them and • The developmental implications and opportunities for how individuals (people and buildings) can enable the health and continuing evolution of place and themselves – through how they function, the qualitative state of being they seek and enable, and what they value and express will toward (adapted from a framework developed by Charles Krone as part of his thinking technology [8]). Capturing the essence understanding that conveys “who” a place is as a living being emerges from the whole systems assessment. Questions used to reveal the essence include: What is at the core of a system, around which it is organized? What is the web or larger context of reciprocal

DEVELOPING THE RIGHT MIND Will

Being

SYSTEMS OF PLACE

Beliefs

Values

Psychological

Cultural

Developmental

Economic

Social

Settlement

Biological

FOODING SHELTERING

Hydrological

VALUE ADDING PROCESSES

Ecological

Geophysical

Function

TRANSACTING ADORNING

RE-CREATING COMMUNING

© Regenesis. Reprinted with permission

Regenerative Development and Design, Fig. 6 Integral Assessment Scope framework. Used as a means of illuminating the core patterns structuring a place

as the basis for “mapping” their dynamic and evolving interrelationships. © Regenesis Group. Reprinted with permission

Regenerative Development and Design

relationships within which it is embedded, since all systems are comprised of smaller systems and part of larger systems? And what is the potential inherent in a living system, since this is the fuel for regeneration – the constant reaching toward being more whole, being more alive? • A simple example of patterns and the essence of a system is offered through a case study of Mahogany Ridge, Idaho, USA [72]. A reductionist approach or an approach that abstracts life into a checklist might state that nothing should be built on existing farmland. This might be a good principle if the agriculture system was truly symbiotic with nature. In this case, waterintensive monoculture farming had nearly destroyed three distinct ecological systems, each of which had played an essential role in shaping the landscape and its potential for life. An integral assessment looked for patterns of life that had enabled high levels of reciprocal relationship between species and ecological niches, patterns that had been obscured and disrupted by farming practices but could be regenerated: The aerial photo in Fig. 6 depicts approximately 3,500 acres of current farmland along the eastern edge of the Big Hole Mountains, just west

135 of the Grand Tetons, that was being considered for development. Originally, these mountain watercourses and alluvial fan supported beaver, otter, native cutthroat trout, salmon, turkeys, grouse, and mega-fauna, such as deer, elk, moose, and bears. These animals were all responsible for carrying nutrients back upstream into the mountains to feed the forest and diversify the terrestrial and riparian ecosystem. Pioneers of European descent arrived in this place 100 years ago and used row-crop agriculture techniques to farm on this alluvial fan. As a result, ninety percent of the water from the Big Hole Mountains (in picture) was being used for agricultural purposes (spray irrigation), the salmon were no longer breeding in the river, the Yellow Tail cutthroat trout were in species decline, the river was polluted from overloads of nitrogen, and the upstream forests were in decline. The area farmers were going out of business or bankrupt due to the short growing season. The farms, in the past, had been used to support local needs. Twenty to forty acre-per-home zoning is planned as the alternative to large farms. Looked at closely, this photo in Fig. 7 reveals that farming was superimposed on top of this alluvial fan between the stream in the mountain valley (top center of the photograph) and the river. The soils mapping indicated in Fig. 8 reveals the pattern more clearly.

Regenerative Development and Design, Fig. 7 Aerial photo of Mahogany Ridge Resort Community site

136

Regenerative Development and Design

Regenerative Development and Design, Fig. 8 Soil map of Mahogany Ridge site showing alluvial fan patterns

Before farming took place here, these radiating streams and drainage ways served as additional corridors of cover for wildlife moving back and forth between the mountains and the river. When farmers settled the land, they diverted this perennial stream along the highest possible course (in elevation) to irrigate fields that were gridded over a highly productive and robust prairie ecosystem. This action severely simplified and destabilized the ecosystem that once was there. The farming pattern did not preserve the integrity of the ecosystem that contained it; rather, this larger healthy pattern was obliterated. The ecological function of this alluvial fan, and one of the core patterns of the ecosystem in this place, is that of a “living bridge” between the mountains to the west and the Teton River. The pattern of a living nutrient bridge between the mountains and the valley that had been revealed in the assessment indicated that a higher level of ecological health could be reestablished in this mountain, alluvial fan, and River system. The development of homes in tight clusters could be used to pay for the restoration of the stream and habitat corridors that originally connected the Teton River and the mountains and provide wildlife corridors as well as many ecosystem services for community residents. To support the reestablishment of wildlife corridors, native grasses would be planted (minimal turf grass), no fences would be allowed, as well as no off-leash dogs to disrupt nesting and the establishment of territory by new wildlife.

By integrating the community into the development and management of these systems, they could produce food (through diversified agriculture and wild harvesting), timber, and other products, as well as the development of a diversified economy while insuring the provision of ecosystem services for their community. The human involvement in these patterns and processes is key to the ongoing regeneration and development of the potential of the site.

Once the essence understanding of a place is developed as a shared context, designing for Harmony with Place, engages the principles of biomimicry, permaculture, and an essential living process framework. The Biomimicry Guild’s Life’s Principles and their Genius of Place program provide guidance and models for establishing locally attuned strategies for design elements through studying the adaptation and survival of local species within the conditions of a particular site and its surroundings (www.biomimicryguild.com). Permaculture principles, which draw both on an understanding of ecology and of how indigenous people engaged with their place, also provides a lens for developing design strategies that respond to site conditions and opportunities in a way that is mutually beneficial.(http://permacultureprinciples. com; www.tagari.com) Malcolm Wells created an

Regenerative Development and Design

environmental checklist for the evaluation of design and development solutions that merges sets of outcomes into a unified whole in A Regeneration-Based Checklist for Design and Construction (www.sbse.org/resources/ docs/wells_checklist_explanation.pdf). The Essential Living Processes framework was developed by Regenesis for setting overarching project aims, goals, and indicators to guide the design and construction process. It is based on the six critical processes that enable living systems to support the evolution of life. They include the ability to support the basis for life processes – nourishment, shelter (habitat), and the generation and exchange of resources for growing and evolving more life. Because humans cannot be separated from today’s living systems, the factors go beyond material factors that form the outer landscape of a place. They also include the “inner landscape” that sources our spirit and will and drives us to cherish and protect the places we inhabit. They include the ability of a living system to create a sense of identity and foster belonging through its culture, to support meaningful and contributory lives, and to invoke the spirit and inspiration that sustains caring. The framework enables setting aims and goals (and later developing indicators and measuring systems) for how the processes generated by the project support ecological, economic, and social health in each of the Regenerative Development and Design, Fig. 9 Framework showing interrelationship of the 6 essential living processes and how they cross ecological, societal, and economic arenas. Used to set holistic, integrative goals and indicators. © Regenesis Group. Reprinted with permission

137

six areas. Figure 9 represents the interrelationship of these processes and the need to integrate and align how they work across ecological, societal, and economic arenas in order to realize wholesystem regenerative effects.

Future Directions While regenerative development and design still occupy a relatively small niche in the larger world of sustainability efforts, interest in regenerative approaches to the built environment is on the rise. Beyond the USA, growing interest has been particularly marked in Australia and New Zealand, including a government commissioned research report that recommended the adoption of regenerative development as a national policy [12]. A number of interrelated factors, working as a system, are creating a favorable climate that is likely to continue to feed interest in regenerative development and design, including: more practitioners encountering the limits of green building approaches to address the global crises; shifting market dynamics and public awareness; the growing influence of the ecological perspective and the ecosystem concept; the movement toward integrative design with its reliance on interdisciplinary teams; and the growing recognition of the need for

138

community engagement and participation to support the behavior changes required for enduring sustainability. In the 1990s, the most discussed issue for aspiring green designers was how to convince clients to incorporate sustainability features. By 2010, the discussions increasingly were about how to meet clients’ demands for making their project “the greenest” of their kind. Over the same period, appreciation and understanding of ecological sustainability and the ecosystem perspective as it applies to human settlements and institutions has been significantly reshaping thinking in such fields as public health, education, economic and community development, and urban planning, as well as design of the built environment. The core concepts of ecological sustainability, especially the concept of seeing communities as ecosystems in which nature and culture, human and natural designed features are interwoven and interdependent, are driving a move toward increasingly systemic and comprehensive goals. These comprehensive goals are in turn defining new standards of sustainability. Projects seeking to be “the greenest” now include social, economic, educational, and aesthetic goals as well as goals around energy efficiency and pollution. More comprehensive goals affecting multiple fields are necessarily stimulating more integrative and interdisciplinary approaches. They are also adding the need to build community support and stewardship to the list of essential design issues. The ecological and ecosystem perspectives are providing a common “language” or set of frameworks across those fields that is facilitating integrative and participatory approaches across disciplines and between design teams and the public, and in the process further reinforcing an ecological worldview. One effect of this system of factors has been the application of explicitly regenerative approaches across a wider spectrum of fields, and the integration of these fields in regenerative design and community development. Regenerative development has already begun to shift the old, building-centric definition of the built environment to include the relationships between and among buildings, infrastructure, and natural systems, as well as the culture, economy, and politics of communities. The

Regenerative Development and Design

concept of place-sourced design that is core to regenerative development is providing a means for engaging the will of a community to align human and natural communities around shared purposes. These shifts are opening up new roles and new challenges for designers. They are being invited to move “from working on things and structures in isolation from their context to the design of living systems with built-in evolutionary capacity,” to use their design skills to foster “the inherent creativity of the systems in which they are working” [10] instead of viewing those systems as a palette for expressing their own creativity. Regenerative development makes possible a new and critically needed role for developers and developments, the full potential of which is still unfolding. In the future, all developments could be called to serve as instruments for reversing ecological damage, enabling ecological evolution, and ensuring economic futures for sustainable livelihoods. We are just beginning to see glimpses of how, through weaving the many stories of place into a mutually appreciative whole, a regenerative development becomes a harmonizing force within communities and among stakeholders, inspiring new relationships to place, and offering new patterns for shaping the fields of development and design.

Bibliography Primary Literature 1. Benyus J (1997) Biomimicry. Harper Collins, New York 2. McDonough W, Braungart M, Anastas PT, Zimmerman JB (2003) Applying the principles of green engineering to cradle-to-cradle green engineering. Environ Sci Technol 37(23):434A–441A 3. McDonough W, Braungart M (2002) Cradle to cradle: remaking the way we make things. Northpoint, New York 4. Orr D (1992) Ecological literacy: education and the transition to a post-modern world. State University of New York Press, Albany 5. Capra F (1996) The web of life: a new scientific understanding of living systems. Anchor Books, New York 6. Tansley AG (1935) The use and abuse of vegetational concepts and terms. Ecology 16:284–307. https://doi. org/10.2307/1930070

Regenerative Development and Design 7. Van der Ryn S, Cowan S (1996) Ecological design. Island Press, Washington, DC 8. Krone C (2001) West coast resource development: session notes. Unpublished transcription of dialogue by members of the Institute for Developmental Processes, Carmel 9. Mollison B (1988) Permaculture: a designers’ manual. Tagari, Tyalgum 10. Mang P, Haggard B (2016) Regenerative development and design: a framework for evolving sustainability. Wiley, Hoboken 11. Hes D, DuPlessis C (2015) Designing for hope: pathways to regenerative sustainability. Routledge, Abingdon, Oxon 12. Jenkin S, Zari MP (2009) Rethinking our built environments: towards a sustainable future. Ministry for the Environment, Manatu Mo Te Taiao, Wellington 13. Lyle JT (1994) Regenerative design for sustainable development. Wiley, Hoboken 14. The Commonwealth (2017) http://thecommonwealth. org/media/news/secretary-general-opens-conference-re generative-development-call-action#sthash.HHbyLUH 8.dpuf. Accessed 1 June 2017 15. Howard SE (2011) Encyclopædia Britannica. http:// www.britannica.com/EBchecked/topic/273428/SirEbenezer-Howard. Accessed 28 June 2017 16. Geddes P (1915) Cities in evolution. Williams & Norgate, London 17. Marcotullio PJ, Boyle G (eds) (2003) Defining an ecosystem approach to urban management and policy development. United Nations University Institute of Advanced Studies (UNU/IAS), Yokohama 18. Pickett STA, Grove JM (2009) Urban ecosystems: what would Tansley do? Urban Ecosyst 12:1–8. Published online: Springer Science+Business Media, LLC 19. Mitsch WJ, Jørgensen SE (1989) Ecological engineering: an introduction to ecotechnology. Wiley, New York 20. Von Bertalanffy L (1968) General system theory: foundations, development, applications. George Braziller, New York 21. Mang P, Reed B (2012) Designing from place: a regenerative framework and methodology. Build Res Inf 40(1):23–38 22. Sanford C (2011) The responsible corporation: reimagining sustainability and sustainability and success. Josey-Bass, San Francisco 23. McHarg IL (1969) Design with nature. Doubleday, Garden City 24. Rodale Institute (2015) Regenerative organic agriculture and climate change: a down-to-earth solution to global warming. Rodale Institute, Kurztown 25. Medard G, Pahl E, Shegda R, Rodale R (1985) Regenerating America: meeting the challenge of building local economies. Rodale, Emmaus 26. Lyle JT (1984) Designing human ecosystems. Wiley, Hoboken 27. SBSE/Wells (1999) www.sbse.org/resources/docs/wells_ checklist_explanation.pdf. Accessed 28 June 2017

139 28. Pliny F (2017) http://cmpbs.org. Accessed 28 June 2017 29. Lyle JT (1993) Urban ecosystems. In Context 35, p43 30. Mang N (2009) Toward a regenerative psychology of urban planning. Saybrook Graduate School and Research Center, San Francisco. http://powersofplace. com/pdfs/Toward_a_Regenerative_Psychology_of_Urb an_Planning.pdf. Accessed 28 June 2017 31. Haggard B, Reed B, Mang P (2006) Regenerative development. Revitalization 1(2):24 32. Haggard B (2002) Green to the power of three. Environ Des Constr 24–31 33. Girardet H (2014) Creating regenerative cities. Routledge, Abington, Oxon 34. Fullerton J (2017) Regenerative capitalism. http:// capitalinstitute.org/regenerative-capitalism. Accessed 8 June 2017 35. Regenerative Business Summit (2017) http:// theregenerativebusinesssummit.com. Accessed 8 June 2017 36. Lush Spring Prize (2017) http://springprize.org. Accessed 8 June 2017 37. The Regenerative Practitioner (2017) http://www. cabal.co.nz/advancing/2017/2/24/the-regenerativepractitioner2017. Accessed 8 June 2017 38. School of Regenerative Design (2017) http:// schoolofregen.org. Accessed 8 June 2017 39. Relief Web (2017) http://reliefweb.int/report/world/ commonwealth-offers-climate-vulnerable-states-lightend-tunnel. Accessed 8 June 2017 40. Cole R, Lorch R (eds) (2012) Special issue: regenerative design and development: current theory and practice. Build Res Inf 40:1 41. Zhang X, Waldron D, de Jong M, Suzuki M, Huisingh D (eds) (2015) Special issue: Toward a regenerative sustainability paradigm for the built environment: from vision to reality. J Clean Prod 109:1–356 42. Colantonio A, Dixon T (2010) Urban regeneration and social sustainability: best practice from European cities. Wiley, New York 43. König A (2013) Regenerative sustainable development of universities and cities: the role of living laboratories. Edward Elgar Publishers, Cheltenham 44. Hemenway T (2015) The permaculture city: regenerative design for urban, suburban, and town resilience. Chelsea Green Publishing, Vermont 45. Wahl D (2016) Designing regenerative cultures. Triarchy Press, Charmouth, Dorset 46. Sanford C (2017) The regenerative business: redesign work, cultivate human potential, achieve extraordinary outcomes. Nicholas Brealey, New York 47. Orr D (1994) Earth in mind. Island Press, Washington, DC 48. Wilson EO (1984) Biophilia. Harvard University Press, Cambridge, MA 49. Heerwagen J (2007) Biophilia and design. Handout for Portland lectures, Aug 2007 50. Zari MP, Storey J (2007) An ecosystem based biomimetic theory for a regenerative built environment. In: Lisbon sustainable building conference, Lisbon

140 51. Kellert S (2004) Beyond LEED: from low environmental impact to restorative environmental design. Keynote address, greening rooftops for sustainable communities conference. Sponsored by Green Roofs for Healthy Cities, Toronto, and City of Portland, Portland, 4 June 2004 52. Anderson MK (2005) Tending the wild: native American knowledge and the management of California’s natural resources. University of California Press, Berkeley 53. Cole R (2010) New context, new responsibilities: building capability, Unpublished paper 54. Williams D (2007) Sustainable design: ecology, architecture, and planning. Wiley, Hoboken 55. Benne B, Mang P (2015) Working regeneratively across scales. J Clean Prod 109:42–52 56. Bailey RG (2002) Ecoregion-based design for sustainability. Springer, New York 57. Reed B (2007) A livings systems approach to design. AIA National Convention May – Theme Keynote address 58. Mang N (2006) The rediscovery of place and our human role within it. Saybrook Graduate School and Research Center, San Francisco. http://powersofplace. com/papers.htm. Accessed 28 June 2017 59. Marvick V, Murphy T (1998) Patterning as process. Permaculture Activist 39:24–27 60. Lerner J (2005) Acupuntura Urbana, Institute for Advanced Architecture of Catalonia, Barcelona 61. Cole RJ, Charest S, Schroeder S (2006) Beyond green: drawing on nature (for the Royal Architectural Institute of Canada’s “Beyond green: adaptive, restorative and regenerative design” course – SDCB 305), The University of British Columbia, Vancouver 62. Mang P (2001) Regenerative design: sustainable design’s coming revolution. Design Intelligence. http://www.di.net/articles/archive/2043/. Accessed 28 June 2017 63. Orr D (2001) Architecture, ecological design, and human ecology. In: Proceedings of the 89th ACSA annual meeting, ACSA, Washington, DC, pp 23–32 64. Ramo J (2009) Age of the unthinkable: why the new world disorder constantly surprises us and what we can do about it. Little Brown and Company, New York 65. von Hayek F (1974) The pretense of knowledge. Nobel Prize acceptance speech 66. Sanford C (2006) Building intelligence: a living systems view. Springhill, Battleground 67. Alexander C (1997) A pattern language: towns, buildings construction. Center for environmental structure series. Oxford University Press, New York 68. Berry W (1981) Solving for pattern in gift of good land. Counterpoint, Berkeley 69. Kauffman S (2008) Reinventing the sacred: a new view of science, reason, and religion. Basic Books, New York 70. Prigogine I (1997) End of certainty. Free Press, New York 71. Urban Sustainability Learning Group (1996) Staying in the game: exploring options for urban sustainability.

Regenerative Development and Design Tides Foundation Project, Center for Neighborhood Technology, Chicago 72. Seven Group, Reed B (2009) The integrative design guide to green building: redefining the practice of sustainability. Wiley, Hoboken

Books and Reviews Aberley D (1994) Futures by design: the practice of ecological planning. New Society Publishers, Gabriola Island Alexander C (2003–2004) The nature of order. The Center for Environmental Structure, Berkeley Bartuska T (1981) Values, architecture and context: the emergence of an ecological approach to architecture and the built environment. In: ACSA annual conference proceedings Bartuska T, Young G (1994) The built environment definition and scope in the built environment: a creative inquiry into design and planning. Crisp Publications, Menlo Park Beatley T, Manning K (1997) The ecology of place: planning for environment, economy, and community. Island Press, Washington, DC Birkeland J (2010) Positive development: from vicious circles to virtuous cycles through built environment design. Earthscan, London Bossel H (2001) Assessing viability and sustainability: a systems-based approach for deriving comprehensive indicator sets. Conserv Ecol 5(2):12. http://www. consecol.org/vol5/iss2/art12/ Charles J, Kibert J (1999) Reshaping the built environment: ecology, ethics, and economics. Island Press, Washington, DC Cowan S (2004) Evaluating wholeness. Resurgence 225:56 Crowe N (1997) Nature and the idea of a man-made world: an investigation into the evolutionary roots of form and order in the built environment. MIT Press, Cambridge, MA Cunningham S (2008) reWealth! McGraw Hill, New York Edwards A (2010) Thriving beyond sustainability: pathways to a resilient society. New Society Publishers, Gabriola Island France RL (ed) (2008) Handbook of regenerative landscape design. CRC Press, Boca Raton Franklin C (1997) Fostering living landscapes. In: Thompson G, Steiner F (eds) Ecological design and planning. Wiley, New York Golley FB (1993) A history of the ecosystem concept in ecology: more than the sum of the parts. Yale University Press, New Haven Graham R (1993) Restorative design: an interview with Bob Berkebile. Designing a sustainable future. In: Context 35:9 Gross M (2010) Ignorance and surprise: science, society, and ecological design. MIT Press, Cambridge, MA Haggard B (2001) The next step: transforming the building industry to model nature. Hope Dance Hawley AH (1950) Human ecology: a theory of community structure. Ronald, New York

Regenerative Development and Design Holling CS (1994) New science and new investments for a sustainable biosphere. In: Jansson A (ed) Investing in natural capital: the ecological economics approach to sustainability. Island, Washington, DC, pp 57–97 Holmgren D (2002) Permaculture: principles and pathways beyond sustainability. Holmgren Design Services, Victoria International Federation of Landscape Architects (2003) Definition of the profession of landscape architecture. IFLA News, no. 48 Kingsland SE (2005) The evolution of American ecology, 1890–2000. Johns Hopkins University Press, Baltimore Lemons J, Westra L, Goodland R (eds) (1998) Ecological sustainability and integrity: concepts and approaches. Kiuwer, Dordrecht Likens GE (1992) The ecosystem approach: its use and abuse. Ecology Institute, Oldendorf/Luhe McDaniel CN (2006) Design on the edge review. http:// homepages.rpi.edu/mcdanc/opEdsAndCommentary/ OrrDesignOnEdgeAnnalsPublished.pdf. Accessed 1 Dec 2017 McIntosh R (1985) The background of ecology: concept and theory. Cambridge University Press, New York McMurry A (2006) Community health and wellness: a socio-ecological approach. Elsevier, Chatswood Melby P, Cathcart T (2002) Regenerative design techniques: practical applications in landscape design. Wiley, Hoboken Nabhan G (1997) Cultures of habitat. Counterpoint, Washington, DC

141 Naess A (1989) Ecology, community and lifestyle: outline of an ecosophy. Cambridge University press, Cambridge Newman P, Jennings I (2009) Cities as sustainable ecosystems: principles and practices. Island Press, London Orr D (2002) The nature of design: ecology, culture, and human intention. Oxford University Press, New York Orr D (2006) Design on the edge: the making of a highperformance building. MIT Press, Cambridge, MA Reed B (2006) Shifting our mental model – “sustainability” to regeneration. Rethinking Sustainable Construction 2006: Next Generation Green Buildings, Sarasota Reed B (2007) Shifting from ‘sustainability’ to regeneration. Build Res Inf 35:674–680 Todd NJ, Todd J (1993) From eco-cities to living machines: principles of ecological design. North Atlantic Books, Berkeley Wahl C (2016) Designing for regenerative cultures. Triarchy Press Ltd., Axminster Wann D (1996) Deep design. Island Press, Washington, DC Zari MP (2007) Biomimetic approaches to architectural design for increased sustainability. Paper number: 033, School of Architecture, Victoria University, Wellington Zimmerman M (2004) Being nature’s mind: indigenous ways of knowing and planetary consciousness. ReVision, Heldref Publications, Gale Group, Farmington Hills Zolli A, Healy AM (2012) Resilience: why things bounce back. Simon and Schuster, New York

Resilient Design Alex Wilson1 and Mary Ann Lazarus2 1 Resilient Design Institute, Brattleboro, VT, USA 2 MALeco, Washington University, St. Louis, MO, USA

Article Outline Glossary [1] Definition of the Subject Introduction The Resilient Design Principles Climate Change Impacts Climate Resilient Design Strategies Resilient Design Planning Process Voluntary Resilience Standards Resilient Design at Different Scales Resilient Design Strategies Resilient Design Strategies for Infrastructure Failures Future Directions Bibliography

Glossary [1] Active systems rely on electricity or fuel to maintain desired conditions. Adaptation The process of adjusting to new conditions in order to reduce risks to valued assets. Adaptive capacity The ability of a person, asset, or system to adjust to a hazard, take advantage of new opportunities, or cope with change. Assets People, resources, ecosystems, infrastructure, and the services they provide. Assets are the tangible and intangible things people or communities value. Climate stressor A condition, event, or trend related to climate variability and change that can exacerbate hazards.

Hazard An event or condition that may cause injury, illness, or death to people or damage to assets. Integrative design An integrative process, interdisciplinary beyond traditional building disciplines, open, and participatory [2]. Mitigation Processes that can reduce the amount and speed of future climate change by reducing emissions of heat-trapping gases or removing them from the atmosphere. Non-climate stressor A change or trend unrelated to climate that can exacerbate hazards. Passive systems Use natural resources such as sunlight, wind, temperature differences, or gravity to achieve desired conditions without the use of electricity or fuel. Risk The potential cost if something of value is damaged or lost considered together with the likelihood of that loss occurring. Risk is often evaluated as the probability of a hazard occurring multiplied by the consequence that would result if it did happen. Sensitivity The degree to which a system, population, or resource is or might be affected by hazards. Vulnerability The propensity or predisposition of assets to be adversely affected by hazards.

Definition of the Subject The Resilient Design Institute [3] has defined resilience as “the capacity to adapt to changing conditions and to maintain or regain functionality and vitality in the face of stress or disturbance. It is the capacity to bounce back after a disturbance or interruption.” Resilient design applies to both existing and new buildings, along with landscapes, neighborhoods, communities, infrastructure, and regions – designed to provide a safer, more livable, better place for people before, during, and after a disrup-

© Springer Science+Business Media, LLC, part of Springer Nature 2020 V. Loftness (ed.), Sustainable Built Environments, https://doi.org/10.1007/978-1-0716-0684-1_1031 Originally published in R. A. Meyers (ed.), Encyclopedia of Sustainability Science and Technology, © Springer Science+Business Media LLC 2018 https://doi.org/10.1007/978-1-4939-2493-6_1031-1

143

144

tion. It is also about adaptation to changing conditions that are resulting from climate change. Resilient design is a companion to sustainable design, but it is different. While many resilient design strategies directly contribute to sustainability (such as more energy-efficient buildings that will maintain livable conditions in the event of an extended power outage), other strategies may conflict with sustainability features, such as redundant power or water systems that have greater embodied energy. This chapter primarily focuses on resilient design recommendations for buildings.

Introduction Resilient design is an emerging priority in the design community. Major storm disruptions in recent years, including Hurricane Katrina in 2005, Tropical Storm Irene in 2011, and Superstorm Sandy in 2012, have drawn attention to the vulnerabilities our buildings face and demonstrated the need for resilience. Resilience became even more front page news after the triple hurricane storms of Harvey, Irma and Maria in summer 2017. As the climate warms over the coming decades and centuries, many of these vulnerabilities are expected to increase. As storms become more intense, sea levels rise, precipitation patterns are altered, and heat waves become more frequent. At the same time, there is now an everpresent risk of terrorist actions and growing concern about cyberterrorism, including a potentially targeted electric grid that could lead to widespread and prolonged power outages [4]. We cannot eliminate these vulnerabilities, but we can design our buildings and communities to be better prepared for these situations and to keep building occupants far safer. This is the purpose of resilient design. This article will examine the key elements of resilient design at the building scale as well as community scale.

The Resilient Design Principles The Resilient Design Principles, developed by the Resilient Design Institute [5], provide a good

Resilient Design

starting point for understanding resilience and resilient design. 1. Resilience transcends scales. Strategies to address resilience apply at scales of individual buildings, communities, and larger regional and ecosystem scales; they also apply at different time scales – from immediate to long term. 2. Resilient systems provide for basic human needs. These include potable water, sanitation, energy, livable conditions (temperature and humidity), lighting, safe air, occupant health, and food; these should be equitably distributed. 3. Diverse and redundant systems are inherently more resilient. More diverse communities, ecosystems, economies, and social systems are better able to respond to interruptions or change, making them inherently more resilient. While sometimes in conflict with efficiency and green building priorities, redundant systems for such needs as electricity, water, and transportation improve resilience. 4. Simple, passive, and flexible systems are more resilient. Passive or manual override systems are more resilient than complex solutions that can break down and require ongoing maintenance. Flexible solutions are able to adapt to changing conditions both in the short and long term. 5. Durability strengthens resilience. Strategies that increase durability enhance resilience. Durability involves not only building practices but also building design (beautiful buildings will be maintained and last longer), infrastructure, and ecosystems. 6. Locally available, renewable, or reclaimed resources are more resilient. Reliance on abundant local resources, such as solar energy, annually replenished groundwater, and local food, provides greater resilience than dependence on nonrenewable resources or resources from far away. 7. Resilience anticipates interruptions and a dynamic future. Adaptation to a changing climate with higher temperatures, more intense storms, sea level rise, flooding,

Resilient Design

drought, and wildfire is a growing necessity, while non-climate-related natural disasters, such as earthquakes and solar flares, and anthropogenic actions, like terrorism and cyberterrorism, also call for resilient design. Responding to change is an opportunity for a wide range of system improvements. 8. Find and promote resilience in nature. Natural systems have evolved to achieve resilience; we can enhance resilience by relying on and applying lessons from nature. Strategies that protect the natural environment enhance resilience for all living systems 9. Social equity and community contribute to resilience. Strong, culturally diverse communities in which people know, respect, and care for each other will fare better during times of stress or disturbance. Social aspects of resilience can be as important as physical responses. 10. Resilience is not absolute. Recognize that incremental steps can be taken and that total resilience in the face of all situations is not possible. Implement what is feasible in the short term, and work to achieve greater resilience in stages.

Climate Change Impacts Climate change is already having observable effects on the environment from increased temperatures, reduced growing seasons, changes in precipitation patterns, more droughts and heat waves, more frequent and intense storms, flooding, and sea level rise [6]. In many cases scientists predict that climate change will make past hazard and climate-related events more frequent and intense. As climate and environmental conditions change, associated negative impacts will as well, so damages from past events can be considered as a baseline when considering what is vulnerable. These impacts can have a cascading impact. For example, drought, which climate models predict will become more frequent and of longer duration in some areas, increases wildfire risk, as does the die-off of trees that can occur when warming winters fail to keep wood-boring beetles in check, as has occurred in the Northern Rockies in recent decades.

145

All these impacts challenge those in the design and construction of the built environment to design future facilities and renovations to adapt to changing yet unclear conditions. With shifts underway, designs should be based on future climate scenarios and conditions rather than rely on historical data. Designing for warming conditions is challenging since future scenarios are unknown at both the macroscale (e.g., business as usual versus different degrees of emission reduction) and the microscale (e.g., understanding local vulnerabilities, risk factors, and program goals). Many new and developing resources are becoming available that define a range of future outcomes to be incorporated into the planning and design process. The impacts of climate change in the United States by region and sector are included in the US National Climate Assessment Report [7]. Many national, international, and regional government resources provide locationspecific climate projections that address temperature, precipitation, sea level rise, flooding, and other attributes. Many cities across the globe are developing climate action plans that identify their high priority climate vulnerabilities and strategies that should be considered in future building and neighborhood design. The Rockefeller Foundation’s 100 Resilient Cities initiative defines a holistic approach to fostering resilience across physical, social, and economic systems in 100 cities across the globe [8].

Climate Resilient Design Strategies Design for climate change requires solutions that are specific to the identified conditions of a particular site and program. Many of the resilient design strategies described in this chapter apply to design for climate change. That being said, there are a few overarching design strategies that particularly apply when considering climate resilience: 1. Seek solutions that promote adaptability and flexibility: Given the many unknown natural and man-made future circumstances, resilient design prioritizes design solutions that enable future adaptations. This includes

146

Resilient Design

strategies such as modular systems or providing room for expansion of systems over time, looking for ideas that can take advantage of change rather than seeking to maintain the status quo. Designing for adaptability and flexibility has the added advantage of potentially minimizing deconstruction and reconstruction, a significant contribution to waste and emissions. 2. Draw solutions and issues from the (future) vernacular: Resilient design looks to vernacular and natural system strategies in climates similar to those projected in your future (see Fig. 1). As with the changes in plant habitat zones, climate science predicts significant shifts in weather patterns so that historical solutions will no longer be valid over time. In Upstate NewYork

1961–1990

2010–2039 2040–2069 2070–2090 2040–2069

2070–2090 Higher-Emissions Lower-Emissions

Resilient Design, Fig. 1 Migrating upper New York state climate (Source: Northeast Climate Impacts Assessment Team (2007) Confronting Climate Change in the US Northeast: Science, Impacts and Solutions, p.7 [11])

addition to studying building solutions from the (future) vernacular, it will be critical to identify strategies based on natural systems in that future climate for inspiration. New threats may need to be addressed due to climate shifts, for example, the spread of insect-borne disease with climate shifts that requires consideration of design solutions to inhibit infiltration [9]. 3. Focus first on the building site and envelope: Resilient design prioritizes solutions that mitigate and adapt to climate shifts in the design of the site and building envelope. Siting and envelope decisions far outlast the building mechanical systems which are on a regular replacement cycle. Solutions include siting buildings and base floor elevations based on future conditions for the anticipated life of the building and providing higher levels of insulation in the building envelope to mitigate against warmer temperatures. In addition to these priority passive solutions, climate shifts will also require active or hybrid systems, strategically managing changes in system loads and incorporating renewables. (See Design for Future Climate [10] by Bill Gething for suggested design strategies.)

Resilient Design Planning Process Integrating resilience into the design and planning process begins at the earliest stages through an integrative design approach. See Fig. 4 as a prototype resilience process diagram. Site- and program-specific variables – such as project location, goals, program requirements and criticality, population exposure, and sensitivity – should determine which hazards and vulnerabilities to address. A resilience design planning process generally includes these steps: 1. Form a team representing a diverse group of stakeholders to review and determine resilience parameters throughout the planning, design, and construction process. The composition of this group will change over time, but core members should be included throughout

Resilient Design

147

to maintain continuity. A resilience standard can help structure the process. 2. Explore and identify hazards related to local weather conditions, climate trends, and natural hazards. Wherever possible, use third-party authoritative resources to identify any hazards that may pose a high risk to the project. Include the potential impact of man-made hazards such as power and water infrastructure failures and terrorism. The team, using the best available information and expertise, should determine their climate projection scenario based on the service life of the facility. Some climate scientists advise that any facility planned for occupancy after 2030 should take into account climate change impacts in order to provide a future habitable, functional space. Figure 2 identifies natural hazards and climate change hazards to be explored. 3. Identify the primary assets, functions, and occupant populations of the facility or set of facilities in your project. In the process of identifying primary assets, it is critical to determine the importance of maintaining functionality in case of disruption and of protecting sensitive populations, such as the elderly, children, or those in healthcare settings. 4. Assess vulnerabilities for the assets, and determine resilience goals. The project team should rate the vulnerability of each primary function and occupancy by considering its Resilient Design, Fig. 2 Common Hazards to Assess in Resilient Planning

sensitivity to disruption based on the identified hazards along with climate and non-climate stressors. Climate stressors exacerbate the impacts of hazards and damages. For example, the Institute of Medicine of the National Academies report on “Climate Change, the Indoor Environment and Health” identified the seriousness of climate change for the indoor environment and occupant health due to shifts in concentrations of indoor pollutants, dampness, infectious agents and pests, and increased thermal stress. Non-climate stressors should also be considered, such as a change in natural drainage due to adjacent planned construction. Once the sensitivity is determined, assess vulnerability of each function by comparing sensitivity against adaptive capacity. Adaptive capacity addresses the ability to adjust to change or improve in new situations. Generally the higher the sensitivity and lower the adaptive capacity, the greater the vulnerability. Finally, determine the risk level for the most vulnerable aspects of the project based on the probability of an occurrence along with its potential damage. Agree on resilience goals and performance outcomes that should be incorporated into the project design based on the assessment process. A spreadsheet available at the US Climate Toolkit website [12] can be used to document the vulnerability and risk assessment process (see Fig. 3).

Natural Hazards to investigate may include but are not limited to: ● Flooding ● Hurricanes ● High-Winds ● Tornado ● Earthquake ● Tsunami ● Wildfire ● Drought ● Landslides and Unstable Soils ● Extreme Heat ● WInter Storms ● Man-made Events Climate Change Hazards may include but are not limited to: ● Sea Level Rise and Storm Surge ● River Flooding ● Temperature Changes ● Precipitation Changes ● Degree Day Changes ● Extreme Weather Events

148

Resilient Design

In column A below, list the most vulnerable or important-toprotect assets you identified on the Vulnerability tab. For each asset, rank the Probability of a Loss as high, medium, or low in column B. In column C, rank the Magnitude of the potential Loss. Once your table is complete, use the rankings to place each asset in the appropriate section of the Risk Charatcterization Matrix.

Probability of a loss >

Risk Characterization Matrix

Magnitude of (potential) loss >

Most Vulnerable Assets

Probability of a Loss (high, med, low)

Magnitude of the Loss (high, med, low) Relative Risk Low Medium High Very High

Resilient Design, Fig. 3 NOAA, risk characterization matrix, US climate resilient toolkit, documenting steps to resilience 170405 Workbook

5. Research and brainstorm potential solutions for the identified priority functions and populations. Use case studies and resilience planning and design resources, such as the Resilient Design Principles, the US Climate Toolkit website, and insights into design strategies further discussed in this chapter. 6. Using an integrative design model, continue to refine the resilience solutions against goals during the entire design and construction process. Develop a plan to monitor the effectiveness of the resilience solutions over the long term, especially when disruptions occur. Make improvements and adaptations as required, and share results (Fig. 4).

Voluntary Resilience Standards As the interest in planning and design for resilience has grown in the past decade, a number of voluntary standards for the built environment are being developed. These standards address resilience at the building, site, community to infrastructure scales, as well as different aspects of resilience: from individual hazards to a broad, holistic analysis. Together, they represent a move toward fostering the integration of resilience into the design of all facilities. The Energy, Kresge, and Barr

Resilient Design, Fig. 4 Los Angeles building resilience planning process. Los Angeles building resilience: a primer for facilities [13]

Foundations completed a study in 2017 summarizing the current state of the US-based resilience standards for facilities (see Fig. 5). Highlights of a few of the building and community-specific standards: • Enterprise Green Communities Strategies for MultiFamily Building Resilience [15]: This

Resilient Design

149

Resilient Design, Fig. 5 Comparative framework of resilience standards. Voluntary resilience standards: an assessment of the emerging market for resilience in the built environment [14], May 2017, page 8

guideline describes 19 design and programmatic strategies to enhance resilience for affordable multifamily housing across four categories: protection, adaptation, backup, and community. In addition to building-related strategies, the guideline includes programmatic concepts to enhance community resilience, outlining a resilience planning process and defining a Green Communities Criteria certification incorporating the strategies. • FORTIFIED standards[16]: Developed by the Insurance Institute for Business and Home Safety, the FORTIFIED standards give specific prescriptive criteria for a system-based

approach to building resilience for extreme weather including hurricanes, high winds and hail, and flooding. The standards apply to commercial properties and include a third-party certification process with multi-tiered certification levels. • LEED Pilot credits: The LEED pilot credits from the USGBC consist of a suite of three credits that cover planning for resilience, design for enhanced resilience, and passive survivability and functionality during disruptions. These are currently under redevelopment. They follow the LEED system with a third-party review to achieve credits.

150

• Resilience-based Earthquake Design Initiative (REDi) [17]: The REDi rating system, developed by Arup, is a standard for resiliencebased earthquake design for buildings. It gives criteria to allow for business operations and livable conditions to resume quickly after a major seismic event. • Resiliency Action List and Credit Catalog (RELi) [18]: RELi is a national consensus standard developed by a consortium of partners that covers buildings, communities, homes, and infrastructure. It incorporates a holistic approach including resilience, restoration, regeneration sustainability, and wellness. The action list is set up with prerequisites and optional credits. There is currently no thirdparty review process. • The National Institute of Standards and Technology (NIST) Community Resilience Planning Guide [19]: A step-by-step guide that includes a triple bottom-line definition of goals to incorporate into economic development, zoning, mitigation, and other planning activities for buildings and infrastructure systems.

Resilient Design at Different Scales Each scale of resilient design provides different opportunities and impacts: Building Scale Resilience New construction projects provide significant opportunities to incorporate resilience goals into decision-making for all aspects of a project from the outset: building siting and massing, landscape design, building layout, envelope design, material selection, and building systems. New construction designs for net zero carbon also address the mitigation goal of reducing greenhouse gas emissions to slow climate change through reducing building sector energy consumption. This is the focus of the Architecture 2030 initiative [20] which promotes zero carbon in the built environment by 2030 for new construction and substantial renovations. Retrofits of the existing building sector, however, provide the greatest near-term impact in fostering resilience. Mitigation can take place

Resilient Design

through large-scale building envelope retrofits and systems upgrades for existing buildings. In addition to mitigation, there is an urgent need to address adaptation of the existing built environment to better prepare for and respond to change through resilient design. Some of these actions can be low cost/no cost and implemented through zoning, building code changes, or recommended best practices, as illustrated by New York City’s Urban Green [21] 33 actionable recommendations developed after Superstorm Sandy, many of which have been put into place. These include actions that address flood and wind hazards (adaptation) as well as energy reductions and transition to renewable sources (mitigation) such as: • • • •

Elevating buildings and building systems Providing sidewalk flood protection Safeguarding toxic material storage Removing barriers to cogeneration and the use of solar energy • Supplying drinking water without power In addition as our planet warms, health scientists predict increased mortality without widespread adaptation measures in existing buildings. According to one study by Columbia University “Towards More Comprehensive Projections of Urban Heat-Related Mortality: Estimates for New York City under Multiple Population, Adaptation, and Climate Scenarios”... [22] “the substantial reduction of heat-related mortality, particularly under the high-adaptation scenario, provides evidence of the importance of public policy measures leading to continuous heat adaptation.” Implementing adaptation actions to address rising temperatures includes insulating and tightening existing building envelopes, increasing solar reflectance and shading, and reducing glazing. These have the mutual benefit of creating more livable environments while gaining emissions reductions.

Neighborhood Scale Resilience Several resilience strategies become more viable when implemented at the neighborhood or larger scale, allowing for closed-loop systems or scaling

Resilient Design

up of impacts. In the study “Avoided Heat-Related Mortality through Climate Adaptation Strategies in Three US Cities,” [23] authors looked at Philadelphia, Atlanta, and Phoenix and found that “combinations of vegetation and albedo enhancement...offset projected increases in heat-related mortality by 40% to 99% across the three metropolitan regions.” Working at the neighborhood scale, building massing and spacing can create self-shading solutions to reduce solar heat gain. When provided at the neighborhood scale, fundamental services increase resilience with less susceptibility to outages and reduced waste: microgrid energy systems, potable water and wastewater, and transit and food systems. Neighborhood services can also add a level of redundancy in case the larger service grid fails in times of disruption. Community Resilience As major disruptions are taking place more frequently in urban centers throughout the world, the recognition of the importance of community-based resilience has grown significantly. As Craig Fugate, former US FEMA administrator, said in 2012, “We found that our first responder is often your neighbor. They are a resource, so you have to figure out how to bring them into the team.” [24] Since then, many government and nonprofit organizations are focused on strengthening community resilience. These efforts work to strengthen community relationships and networks, expand local knowledge, and improve neighborhood connections to services. The importance of community resilience reinforces the role that specific building types can play toward this goal. For example, local schools, libraries, and community centers can serve the dual role of providing neighborhood gathering places for ongoing community building while acting as accessible resilience refuges or cooling centers in times of disruption. Other “third places” such as local coffee shops could also serve in this capacity.

Resilient Design Strategies Resilient design involves a collection of strategies that are designed to protect buildings and keep building occupants safe from natural disasters or

151

other events, including terrorism and accidental power outages. Such strategies are highly variable and depend on the location, building type, and building design – as explained in the sections that follow: Wind-Resistant Construction Practices As climate change advances over the coming decades, the intensity of storms is expected to increase. According to the Geophysical Fluid Dynamics Laboratory of the National Oceanic and Atmospheric Administration at Princeton University, warmer temperatures in the Atlantic Ocean and Gulf of Mexico will result in warmer air temperatures and more intense tropical storms [25]. At the same time, warmer land temperatures in the United States are expected to increase convective air potential energy (CAPE). According to the NASA Earth Observatory [26], CAPE is a measure of how much raw energy is available for storms – the cause of severe thunderstorms, supercells, derechos, and tornadoes. As a result, intense storms with high winds are expected to become a greater concern in the decades ahead. Buildings that are able to withstand high winds from hurricanes, severe thunderstorms, supercells, and derechos are built incorporating best practices for wind, such as specialized building codes in hurricane regions. The Resilient Design Institute proposes that the Miami-Dade County Hurricane Code, or a similar standard, be adopted nationwide or used as a guideline for structural reinforcement [27]. Wind-resistant construction practices vary widely depending on the building type. With light-frame construction, FEMA provides in-depth guidance on construction practices in high-wind locations, including volume 2 of the Coastal Construction Manual [28]. Another useful resource, also for light-frame construction, is the FLASH (Federal Alliance for Safe Homes) Resilient Design Guide [29]. Except for certain emergency services buildings, it is not economically feasible to design buildings to withstand tornadoes, which can have winds exceeding 250 miles per hour. For buildings in the most tornado-prone locations, it makes more sense to provide safe rooms, where

152

building occupants can safely ride out tornadoes. Even in EF5 tornadoes, the most severe, there are no known examples where people were killed when sheltered in a property designed safe room. Guidelines for safe room construction are available from FEMA [30]. Design for Flooding Flooding is projected to become a more significant problem with our changing climate: sea level rise will make coastal flooding more frequent and lead to tidal flooding even in clear weather; more intense tropical storms will cause more frequent and more severe coastal flooding; more intense rainfall events will contribute to more frequent and more severe inland flooding. Given these vulnerabilities, design for flooding is one of the most important resilient design priorities. Key aspects on flood-resilient design are the following: 1. Avoid flood-prone building sites. While FEMA and NFPI guidelines generally call for no construction in the 100-year flood zone (1% risk of flooding each year), many experts recommend significantly exceeding that minimum, by avoiding the 500-year flood zone (0.2% risk of flooding each year) or keeping buildings well above the base flood elevation (BFE). The LEED pilot credits on Resilient Design call for keeping occupied spaces above BFE + 5 (5 feet above the BFE). 2. Within the 500-year flood zone, incorporate wet floodproofing practices in residential buildings. With wet floodproofing, floodwater is allowed to enter and exit a building through openings in the foundation or lower floors so that hydrostatic forces do not destroy the building or result in significant structural damage. 3. Within the 500-year flood zone, incorporate dry floodproofing practices in new commercial buildings. With dry floodproofing, water is kept out of the building using flood barriers, floodgates, and other components. Significant hydrostatic forces will result on foundation and well systems, so this technique requires careful structural engineering and is usually not feasible for older buildings.

Resilient Design

4. All flood-prone locations in a building – and even in buildings that are not located in flood zones as a defense against flooding caused by leaking pipes or the roof – use building materials and components that can dry out after flooding without causing permanent damage or risk of mold growth. In general, cellulosic materials should be avoided in flood-prone locations in a building, including paper-faced drywall, wood, and carpeting. 5. Keep mechanical and electrical equipment out of any floodable areas in a building. In existing buildings that are in flood zones, mechanical and electrical equipment should be elevated in the building. In commercial buildings, locating mechanical systems on the roof ensures they will not be subject to flooding. Design for Earthquakes Seismic risk is a function of the underlying geology and location of tectonic plates. Most areas with significant seismic risk in the United States are well known and mapped by the US Geological Survey [31]. In the Lower 48 States, the greatest areas of risk are the Pacific Coast, the Western Mountain States, and several locations in the Midwest and Southeast. New understanding of the Cascadia Subduction Zone fault off the coast of Oregon and Washington has elevated awareness of risk in those states in recent years [32]. There has also been an increase in seismic activity in places with extensive oil and gas development such as Oklahoma. The “induced earthquake” activity is likely a result of the wastewater injection that occurs with well drilling and hydraulic fracturing (fracking) for oil and gas recovery [33]. Minimizing earthquake risk is in part about site selection: avoiding sites on or close to earthquake faults. But because earthquakes extend over wide areas – far from the actual faults – building design to minimize earthquake damage is also critically important. Seismic codes are an important part of earthquake design. Such codes have been an important part of building codes since the Great San Francisco Earthquake in 1906, but it is important to understand that the intent of seismic codes is to

Resilient Design

provide for safe egress from a building during and immediately after an earthquake. Building to these codes is not a guarantee that the building will maintain functionality after an even modest earthquake. For this reason, there has been an interest in earthquake resilience standards that go beyond standard seismic codes. The most robust of these is the REDi Rating System developed by seismic engineers at the international engineering firm Arup. A new San Francisco building, 181 Fremont, completed in 2016, is the first large building to utilize the REDi Rating System in its design. The developer recognized the benefit of maintaining functionality in the event of a major earthquake and invested the extra cost in measures needed to achieve the rating. The building’s insurer also recognized the benefits of REDi certification and discounted the insurance by $100,000 per year [34]. Design for Landslides As land to build on becomes scarce, developers are turning to more marginal land, including building lots on steep slopes. Depending on the soils, buildings on such sites may be vulnerable to landslides. Furthermore, with climate change, precipitation patterns are expected to change, with more intense storms and more rapid melting of winter snowpack causing more severe flooding, which may exacerbate landslides on steeper slopes. Designing for landslides is mostly about avoiding steep or unstable sites. In hilly terrain, collaboration with a civil engineer, soil scientist, or geotechnical expert can identify landslide risk. Local universities with a geology department can be a good source of information on landslide risk. Maps showing historical landslides are also valuable because landslides tend to repeat. The US Geological Survey (USGS) maintains a Landslide Hazards Program, offering a wealth of information on landslide vulnerability, warning signs of landslides, and what to do in the event of a landslide [35]. Their Landslide Overview Map of the Conterminous United States [36] provides a general overview of landslide risk, but the map is dated (1982) and the resolution inadequate

153

to inform siting decisions on a local level. More detailed information on landslide risk is often available from state and local governments. Many states, particularly in the West, maintain state geological surveys and publish geologic hazard maps. On sites where landslide vulnerability is a concern, there are measures that can be taken to protect buildings, such as maintaining existing vegetative cover on the land, providing deep footings designed by engineers expert in landslide risk and mitigation, and modifying surface runoff and mudflows through the construction of swales and channels. It also makes sense to provide flexible pipe fittings for buried natural gas and water lines to reduce risk of breakage or leaks. Design for Extreme Heat According to the US EPA, “Extreme Heat Event (EHE) conditions are defined by summertime weather that is substantially hotter and/or more humid than average for a location at that time of year. Because how hot it feels depends on the interaction of multiple meteorological variables (e.g., temperature, humidity, cloud cover), EHE criteria typically shift by location and time of year” [37]. Concerns are greatest when the number of days lengthen to weeks or more. Their effects are particularly significant in urban settings due to the urban heat island effect which can lead to increased mortality, particularly for vulnerable groups like the very young, the elderly, and disadvantaged populations without access to air conditioning. The 2003 heat wave in Europe is estimated to have caused 30,000 deaths [38]. The impact was especially strong because the existing built infrastructure of the typically mild summer climate could not adapt to the extreme heat. There are some basic strategies that should be followed in the design for extreme heat and future warming conditions. Many of these are consistent with fundamental passive design solutions found in much further detail elsewhere in this Encyclopedia. A useful resource is “Design for Future Climates” by Bill Gething which outlines three major approaches to design for extreme heat: on

154

the building site, reduce the heat island effect and provide for external spaces by incorporating maximizing shade, native planting, and permeable pavement systems. For the building enclosure, design to reduce heat gain and allow for passive cooling, incorporting improved envelopes, roof coverings that meet the Cool Roofing Rating Council [39] standards, shading devices and natural ventilation systems. For the building conditioning systems, design for optimum cooling efficiency and future expansion and flexibility. If not currently available, systems should be designed to allow for future connections to renewable energy and/or district energy solutions. Design for Drought Climate change is predicted to increase precipitation in some regions and seasons and reduce it in others. In the mid-latitudes, in particular, increased summer drying is expected [40]. With drought, the output of hydroelectric generators drops, and with severe drought, thermoelectric power plants can be affected, as the water level in reservoirs and other surface waters drop and cooling water becomes less available. (The vast majority of US power plants are thermoelectric plants that depend on cooling water for their operation.) Severe drought can also result in infrastructure failures that contribute to wasted water and water shortages. For example, during the severe drought in Fort Worth, Texas, in 2011, shrinking clay soils resulted in more than 200 water main breaks. The same year, drought-related wildfires in Lubbock, Texas, knocked out power to 20% of the city’s water supply for 2 weeks [41]. Creating buildings that are more resilient to drought involves, first and foremost, maximizing water efficiency. Specifying the most waterconserving plumbing fixtures and appliances goes a long way toward reducing water consumption and allowing continued functioning with curtailed water supply. Outdoor landscapes should be designed for xeriscaping using native, drought-adapted plantings that can survive without irrigation. Beyond water conservation measures, we can rely on on-site water storage. Water storage can be

Resilient Design

integrated with rainwater harvesting, or it can rely on municipal or well water. Cisterns are available in a wide range of sizes and can be buried, ground mounted, or roof mounted. With elevated cisterns (rooftop or otherwise), delivery of water by gravity flow can enhance resilience. To be used for potable applications, harvested rainwater or water that has been stored for any length of time in a cistern should be filtered and treated. However, a cistern designed for landscape irrigation without a high level of filtration can be used in emergency situations with a simple ultrafine hand filter or treatment with water treatment tablets to kill bacteria and viruses. Redundant water supplies can further enhance resilience. Buildings on municipal water systems would be more resilient if they also had rainwater harvesting systems and/or a drilled well. Rural homeowners with deep wells and submersible pumps might want to develop a spring (if feasible) or install a rainwater harvesting system. Design for Winter Storms Winter storms have for centuries brought hardship to affected communities, and that is expected to continue even as the planet warms overall. The changing precipitation patterns described above may also increase winter storm intensity. In more northern climates, being prepared for winter storms is an important component of resilient design. Preparing for winter storms involves both building design and a strategy for ensuring safety and functionality during power outages. Relative to building design, preparing for winter storms may involve the following: • Ensuring adequate structural design to carry significant snow loads. • Ensuring high levels of insulation and air tightness to capture internal heat sources. • Designing an air-sealing and/or roofventilation strategy to minimize the risk of ice dams. • Ensuring proper design to minimize the risk of frozen pipes, such as avoiding running water pipes through exterior walls.

Resilient Design

• Providing a plan for removing snow during extreme snowfall events. This might include stair access to low-slope roofs on commercial buildings and space on the building site to store snow piled up by plowing. • Ensure building safety and functionality in the face of power, gas, and water infrastructure failures, to be further discussed below.

Resilient Design Strategies for Infrastructure Failures Design for Power Outages: Passive Survivability Extended power outages may occur as a cascading impact of many natural disasters as well as certain terrorist actions (including cyberterrorism). For these reasons, we should be designing buildings – particularly homes, apartment buildings, schools, and other buildings that may serve as shelters – so that they will maintain habitable temperatures should they lose power. This is the concept of passive survivability, an idea first proposed by BuildingGreen [42] and addressed in The New Orleans Principles [43]. To achieve passive survivability (or thermal habitability), the following design features should be included: 1. A highly insulated building envelope. 2. High-performance windows or glazing systems (triple-glazing and dual low-emissivity coatings recommended in colder climates). 3. Tight construction – achieving Passive House standards for air tightness is recommended, with mechanical ventilation relied on during normal building operation to ensure adequate fresh air. 4. Cooling-load avoidance measures, including such features as overhangs above windows, vegetative shading, window treatments to control solar gain (exterior roller shutters, interior insulating blinds, interior louvered shutters, reflective coatings, etc.). 5. Passive solar heating with more glass oriented to the south in all climates with heating loads and thermal mass to store heat.

155

6. Natural ventilation. 7. Daylighting Backup and Redundant Power Systems While buildings should be able to maintain habitable temperatures during power outages (passive survivability), a greater level of functionality can be retained in a building that has backup power. Backup power systems come in various forms and provide varying levels of automation and controls. The most common options are described below: Portable, gasoline-fuel generators. The most prevalent backup power systems for homes are fuel-fired generators. Portable, gasoline-fueled generators are common for homes, with outputs ranging from about 500 watts to about 10 kilowatts (kW). These can be started up outdoors, and critical electrical loads can be plugged into them: emergency lighting, a refrigerator, radio, etc. Most of these portable generators are not wired into the home’s electrical panel or even a critical-loads subpanel. Most are totally separate from the house wiring and require the homeowner to plug the appliance, lighting fixture, and other electric loads into outlets on the generator. Larger generators have more outlets than smaller ones, and some offer 220 volt as well as 110 volt output. For safety reasons, they should be operated outdoors, not inside a building or even a garage. Stationary generators. Stationary generators are typically larger than portable generators. These are typically fueled by diesel, propane, or natural gas, and they range in size from about 5 kW to several hundred kW, with much larger generators serving large commercial and industrial facilities. Access to Potable Water For situations in which building occupants must shelter in place following a disaster, access to potable water is critically important. How this is addressed depends on the water source for the building. According to the USGS, approximately 86% of the US population lives in buildings served by municipal wastewater systems; the remaining 14% are on private wells, springs, or other sources [44].

156

Most municipal water systems are reasonably resilient. The pumps needed to deliver the filtered water typically have backup generators to maintain operation during power outages. So, unless the water mains are damaged or there are problems at the pumping station or filtration plant (as could happen with an earthquake), municipal delivery of potable water to the building can usually be assumed. In high-rise buildings served by municipal water, however, there are often pumps within the building that are needed to deliver water to higher floors. In New York City, for example, gravity flow from reservoirs serving the city will deliver water to about the 6th floor, but pumps are needed within the buildings to deliver water to higher floors (and provide adequate water pressure for even low-rise buildings) [45]. To address the need to provide residents of taller buildings with access to potable water, faucets can be installed on lower floors (low enough that the municipal water is delivered without in-building pumps) that are accessible to all residents. This solution is now required in New York City for all new high-rise residential buildings [46]. To protect against the prospect of a municipal water source becoming inoperable or unavailable as a result of a disaster or other event, potable water can be stored onsite. The Sphere Handbook, an internationally recognized set of “common principles and universal minimum standards in lifesaving areas of humanitarian response,” lists per capita potable water requirements as follows [47]: • Survival needs (water intake): 2.5–3 liters per day (0.7–0.8 gallons per day), depending on climate and individual physiology • Basic hygiene practices: 2–6 liters per day (0.5–1.6 gallons per day), depending on social and cultural norms – although this does not include toilet flushing • Cooking: 3–6 liters per day (0.8–1.6 gallons per day), depending on food type and social and cultural norms • Totals: 7.5–15 liters per person per day (2.0–4.0 gallons per person per day) The duration of an emergency for which water should be stored is hard to predict. For residential

Resilient Design

buildings in earthquake-prone locations, adequate water storage for 5 or 7 days is reasonable. Water can be stored in packaged water bottles, 5-gallon carboys, specialized vessels for water storage in buildings, or more sophisticated cisterns. Keeping water containers sealed to keep out contaminants is advised, as is keeping them in a dark location to prevent algae growth. For buildings that are not served by a municipal water supply, resilience can be provided either by storing water (as above) or by providing an alternative means of pumping or delivering water. Most buildings with their own water supplies are rural homes that are served by deep wells and submersible pumps. Emergency water supplies can be provided by having a backup generator to power the submersible pump (note that most submersible pumps in North America are 240 volt), a solarpowered pump that can work during the day on sunny days, or a hand pump. State-of-the-art hand pumps today maintain their prime for months, include weep holes for freeze protection, and can pump from a depth of several hundred feet [48]. In rural areas, some homeowners may have access to a spring. If that is an option, the spring should be tested to ensure that the water quality is adequate. This provides the ultimate in water resilience. Resilient Wastewater Systems Depending on the building and the toilet fixtures, wastewater disposal can be one of the most challenging problems in the event of a disaster, power outage, or loss of municipal wastewater. Gravity-flush toilets (with tanks and removable lids) tend to be more resilient than “blow-out” toilets that rely on water pressure for flushing. If a power outage eliminates the water source or there is a problem with the water supply, the toilet tank of gravity-flush toilets can be manually filled with water by removing the lid and then flushed as usual. That is not possible with blow-out toilets relying on water pressure for flushing or for pressure-assist toilets that have an inner pressure tank within the tank above the toilet bowl. There may also be a problem with toilet flushing in multifamily residential or commercial buildings when the sewer line in the basement has been

Resilient Design

closed off with a manual valve to protect against backflow into the building through the sewer line. Backflow preventers (one-way valves) should not be a problem, but manual-closure valves are sometimes installed as an extra precaution. The most resilient wastewater system is a composting toilet system with a large composting chamber located beneath the toilet – usually in the basement. Most composting toilets do not require water or electricity (other than for ventilation). An exception is a composting toilet served by foamflush toilets; these require both electricity and a small amount of water for operation. Usually, the compost vessel in a commercially sized composting toilet will not fill up following a disaster, because most interruptions are short term, but in an exceptional climate or infrastructure failure event, emptying out waste before that waste is fully composted may be necessary. In such a situation, significant caution should be exercised, because pathogens from the human waste could be present. In larger multifamily buildings, the suggestion is sometimes made to provide two composting toilets (one for men, one for women) on the ground floor that can be used in the event of an emergency [49].

Future Directions We live in a world of complacency. Most of us have enough on our plates that we do not want to add new concerns – like what to do about a potential flood or earthquake or power outage due to cyberterrorism. That complacency is interrupted, from time to time, by major events, such as Hurricane Katrina, Superstorm Sandy, the California Drought of the 2010s, and “Snowmageddon” in February 2010. Immediately after such events – sometimes for several years – there is a lot of interest in how to enhance our resilience to climate disturbances. But as memories fade, that awareness of the need for greater resilience gradually shifts to the back burner to make room for other concerns. In some cases, our complacency is interrupted by predictions about potential events – as

157

occurred in the Pacific Northwest following publication in The New Yorker of an article, “The Really Big One,” with the attention-grabbing lead: “An earthquake will destroy a sizeable portion of the Pacific Northwest. The question is when” [50]. That article on the Cascadia Subduction Zone fault inspired a flurry of resilience actions in Portland and other cities in Oregon and Washington. But such instances are fairly rare. We tend to sit back and then express surprise when something happens, or we blame officials for not doing something about that risk. Each of these events, however, does move the needle forward, even if slightly. We have been lucky, to date, that the well-publicized, widespread power outages that affected major population centers of the US over the past several decades have occurred during moderate weather. A week-long power outage in New York City in July, with temperatures in the upper 1990s and relative humidity over 90%, would likely result in significant fatalities and increase demand for changing building codes to ensure that at least new buildings provide for passive survivability. Climate change and global terrorism are likely to increase the frequency of these crises, reminding us that resilient design is a high priority. Resilience will move back onto the front burner, and we will see many resilient design initiatives and codes implemented. We should take action to avoid the losses and hardships that accompany the increasing number of natural and man-made disasters, by ensuring that resilient design becomes standard design.

Bibliography Primary Literature 1. U.S. Climate Resilience Toolkit. Glossary. https:// toolkit.climate.gov/content/glossary 2. Group R (2009) The integrative design guide to green building: redefining the practice of sustainability. Wiley, Hoboken 3. Resilient Design Institute. http://www.resilientdesign. org/what-is-resilience/ 4. Koppel T (2015) Lights out: a cyberattack, a nation unprepared, surviving the aftermath. Crown, New York 5. Resilient Design Institute. http://www.resilientdesign. org/the-resilient-design-principles/

158 6. NASA Global Climate Change. The consequences of climate change. https://climate.nasa.gov/effects 7. National Climate Assessment Report (2014). http:// nca2014.globalchange.gov/report 8. Resilient Cities. http://www.100resilientcities.org/ 9. Institute of Medicine; Board on Population Health and Public Health Practice; Committee on the Effect of Climate Change on Indoor Air Quality and Public Health (2011) Climate change, the indoor environment and health. https://www.nap.edu/catalog/13115/ climate-change-the-indoor-environment-and-health 10. Gething B (2010) Design for future climate, technology strategy board. http://www.arcc-network.org.uk/ wp-content/D4FC/01_Design-for-Future-Climate-BillGething-report.pdf 11. Northeast Climate Impacts Assessment Team (2007) Confronting climate change in the U.S. Northeast: science, impacts and solutions, p 7. http://www. ucsusa.org/sites/default/files/legacy/assets/documents/ global_warming/pdf/confronting-climate-change-in-theu-s-northeast.pdf 12. NOAA. U.S. Climate Resilience Toolkit. https:// toolkit.climate.gov/steps-to-resilience/overview 13. USGBC (2016) Los Angeles Chapter, Building Resilience Los Angeles: A Primer for Facilities, p iii. http:// www.resilience.la/#intro 14. Meister Consultants Group (2017) Voluntary resilience standards: an assessment of the emerging market for resilience in the built environment, report for the energy, Kresge and Barr foundations, p 8. http://www. mc-group.com/wp-content/uploads/2017/05/MCGVoluntary-Resilience-Standards-Report.pdf 15. Enterprise Green Communities (2015) Ready to respond: strategies for multifamily building resilience. http://www.enterprisecommunity.org/resources/readyrespond-strategies-multifamily-building-resilience13356 16. Insurance Institute for Business & Home Safety. FORTIFIED standards. https://disastersafety.org/ibhs-businessprotection/fortified-new-ibhs-commercial-constructiontools-protect-the-bottom-line/ 17. Arup. REDi™ rating system. http://publications.arup. com/publications/r/redi_rating_system 18. The RELi Collaborative. RELi resilience action list & credit catalog. http://c3livingdesign.org/?page_id=5110 19. National Institute of Standards and Technology. U.S. Department of Commerce, Community resilience planning guide. https://www.nist.gov/topics/communityresilience/community-resilience-planning-guide 20. Architecture2030. http://architecture2030.org/ 21. Urban Green (2013) Building resiliency task force. http://urbangreencouncil.org/content/projects/buildingresilency-task-force 22. Petkova E, Vink J, Radley M, Gasparrini A, Bader D, Francis J, Kinney P (2017) Towards more comprehensive projections of urban heat-related mortality: estimates for New York City under multiple population, adaptation, and climate scenarios. Environ Health Perspect. https://doi.org/10.1289/EHP166. https:// ehp.niehs.nih.gov/ehp166/

Resilient Design 23. Stone B Jr, Vargo J, Liu P, Habeeb D, DeLucia A, Trail M et al (2014) Avoided heat-related mortality through climate adaptation strategies in three US cities. PLoS One 9(6):e100852. https://doi.org/10.1371/journal. pone.0100852 24. Your First Responder in a Natural Disaster is Often Your Neighbor Says Craig Fugate (2012) http://www. mccno.com/your-first-responder-in-a-natural-disasteris-often-your-neighbor-says-craig-fugate/ 25. NOAA Geophysical Fluid Dynamics Laboratory. Princeton University, Global Warming and Hurricanes. https://www.gfdl.noaa.gov/global-warmingand-hurricanes/ 26. NASA Earth Observatory (2013) Global climate change: vital signs of the planet. Severe thunderstorms and climate change. https://climate.nasa.gov/news/ 897/severe-thunderstorms-and-climate-change/ 27. Resilient Design Institute (2012) Fundamentals of resilient design #2: designing homes for more intense storms. http://www.resilientdesign.org/designinghomes-for-more-intense-storms/ 28. Federal Emergency Management Agency (2011) Guidance document P-55. Coastal construction manual, vol 2, 4th edn. https://www.fema.gov/medialibrary/assets/documents/3293 29. Federal Alliance for Safe Homes (FLASH). Resilient design guide: high wind wood frame edition. flash.org/ resilientdesignguide.pdf 30. Federal Emergency Management Agency. Refer to guidance documents P-320, Taking shelter from the storm: building a safe room for your home or small business, 4th edn, Dec 2014, and P361, Safe rooms for tornadoes and hurricanes: guidance for community and residential safe rooms, 3rd edn, Mar 2015. https://www.fema.gov/safe-rooms 31. USGS Earthquake hazards program – seismic hazard maps and site-specific data. https://earthquake.usgs. gov/hazards/hazmaps/ 32. Pacific Northwest Seismic Network. https://pnsn.org/ outreach/earthquakesources/csz 33. USGS Earthquake hazards program – induced earthquakes. https://earthquake.usgs.gov/research/induced/ myths.php 34. Personal communication with Ibrahim Almufti, P.E., Arup 35. USGS Landslide hazards program. https://landslides. usgs.gov/learn/prepare.php 36. USGS Landslide overview map of the conterminous United States. https://landslides.usgs.gov/hazards/nationalmap/ 37. US EPA 430-B-16-001 (2006) Updated Appendix A (2016) Excessive heat events guide. p 9. https:// www.epa.gov/sites/production/files/2016-03/documents/ eheguide_final.pdf 38. United Nations Environmental Programme (2004) Impacts of summer heat wave in Europe. http://www.unisdr.org/files/1145_ewheatwave.en.pdf 39. Cool Roof Rating Council. http://coolroofs.org/ 40. IPCC Fourth Assessment Report (2007) 9.5.4.2 – Global precipitation changes. https://www.ipcc.ch/ publications_and_data/ar4/wg1/en/ch9s9-5-4-2.html

Resilient Design 41. Wilson A (2012) Fundamentals of resilient design: water in a drought-prone era. Resilient design institute. http://www.resilientdesign.org/fundamentalsof-resilient-design-8-water-in-a-drought-prone-era/ 42. BuildingGreen Report. https://www.buildinggreen. com/op-ed/passive-survivability 43. U.S. Green Building Council. http://www.usgbc. org/resources/new-orleans-principles-celebrating-richhistory-new-orleans-through-commitment-sustainable 44. USGS Water Science School. https://water.usgs.gov/ edu/wups.html 45. Urban Green Council (2013) Building resiliency task force full proposals. http://www.urbangreencouncil. org/content/projects/building-resilency-task-force 46. Local Law 110 (2013) To amend the New York city plumbing code and the administrative code of the city of New York, in relation to requiring residential buildings to provide drinking water to a common area supplied directly through pressure in the public water main. Urban green council: building resiliency task force proposal tracker. http://www. urbangreencouncil.org/resiliencytracker 47. The Sphere Project (2011) The sphere handbook: humanitarian charter and minimum standards in humanitarian response, 3rd edn. http://www. sphereproject.org/handbook/ 48. Wilson A (2012) Hand pumps: an option for backup water pumping. Resilient Design Institute. http:// www.resilientdesign.org/hand-pumps-an-option-forback-up-water-pumping/ 49. Wilson A (2006) Passive survivability: a checklist for action, accompanying the article passive survivability: a new design criterion for buildings. BuildingGreen report. https://www.buildinggreen.com/feature/ passive-survivability-new-design-criterion-buildings/ checklist/1 50. Schulz K (2015) The really big one. The New Yorker. http://www.newyorker.com/magazine/2015/07/20/thereally-big-one

159

Books and Reviews

Adams M, Watson D (2011) Design for flooding: architecture, landscape, and urban design for resilience to climate change. Wiley, Hoboken Enterprise Community Partners (2015) Strategies for multifamily housing resilience. Enterprise Community Partners, New York Eskew-Dumez-Ripple (2014) A framework for resilient design. EDR, New Orleans Gething B (2010) Design for future climate. Technology Strategy Board (Innovate UK), London Hyde R (2000) Climate responsive design. E & FN Spon, New York Koppel T (2015) Lights out: a cyberattack, a nation unprepared, surviving the aftermath. Crown, New York Linnean Solutions, Built Environment Coalition, Resilient Design Institute (2013) Building resilience in Boston: best practices for climate change adaptation and resilience for existing buildings. Boston Society of Architects, Boston Maclay W (2014) The new net zero. Chelsea Green, White River Junction Mcgregor A, Roberts C, Cousins F (2013) Two degrees: the built environment and our changing climate. Routledge, London Resilient Design Institute (2012) Resilient design principles. www.resilientdesign.org Roaf S, Crichton D, Nicol F (2009) Adapting buildings and cities for climate change: a 21st century survival guide, 2nd edn. Routledge, London Rodin J (2014) The resilience dividend. PublicAffairs, New York Urban Green (2013) Building resiliency task force: full proposals. Urban Green, New York

Part II Urban Design for Sustainability

Sustainable Urbanism Douglas Farr Founding Principal, Farr Associates, Chicago, IL, USA

Article Outline Glossary Definition of the Subject Introduction Topic 1: Building a Movement Topic 2: Thresholds of Sustainable Urbanism Future Directions Bibliography

Glossary Biophilia The human love of nature based inon the intrinsic interdependence between humans and other living systems. Charrette A planning session in which participants brainstorm and visualize solutions to a design issue. Comprehensive plan (also master plan or general plan) A municipal document(s) based upon establishing long-term goals and objectives that serves as a guide for making land use changes, preparation of capital improvement programs, and the rate, timing, and location of future development. Elements include economic development, environment, housing, land use, recreation and open space, and transportation. Congress for the New Urbanism (CNU) An organization founded in 1993 by six architects – Peter Calthorpe, Andrés Duany, Elizabeth Moule, Elizabeth Plater-Zyberk, Stefanos Polyzoides, and Daniel Solomon – united around a shared vision of promoting traditional urbanism as an antidote to conventional sprawl created an ad hoc organization to convene annual congresses.

Degree-day A measurement of indoor heating requirements affected by outside temperatures. The number of degree-days for any given day is calculated by subtracting the mean outside temperature from 65  F, and the total degreedays for any longer period is the sum of the degree-days of the individual days in that period. Green building or green design Building design that yields environmental benefits, such as savings in energy, building materials, water consumption, and reduced waste generation. Greenfield Newly developed real estate on what was previously undeveloped open space. Infill development Development occurring within established areas of a city. Integrated design A design approach that optimizes the performance of a building as an entire system, which improves a building’s performance at little or no added cost simply by shifting money within the project. LEED-ND USGBC’s Leadership in Energy and Environmental Design (LEED) standards for Neighborhood Development (ND), which expands the LEED brand for green buildings beyond the scale of the individual building to address multiple buildings, infrastructure, and neighborhood-scale developments. Mixed-use A development that combines residential (multifamily such as fourplexes, condos, apartments, townhouses), commercial, retail, and/or office uses, either in a vertical fashion (i.e., a single building that allows residential uses above ground-floor commercial uses) or a horizontal fashion (i.e., a neighborhood urban center of adjacent buildings and uses clustered together in a development less than 40 acres). New urbanism Neighborhood design approach to promote community and livability. Characteristics include narrow streets, wide sidewalks, porches, and homes located closer together than typical suburban designs.

© Springer Science+Business Media, LLC, part of Springer Nature 2020 V. Loftness (ed.), Sustainable Built Environments, https://doi.org/10.1007/978-1-0716-0684-1_1044 Originally published in R. A. Meyers (ed.), Encyclopedia of Sustainability Science and Technology, Springer Science+Business Media LLC 2018 https://doi.org/10.1007/978-1-4939-2493-6_1044-1

163

164

Planning The process of setting development goals and policy, gathering and evaluating information, and developing alternatives for future actions based on the evaluation of the information. Sprawl Patterns of urban growth that include large acreage of low-density residential development, rigid separation between residential and commercial uses (and/or walled residential subdivisions that do not connect), leapfrog development (areas of vacant land between areas of developed land), minimal support for non-motorized and non-automobile transportation methods, and a lack of integrated transportation and land use planning. Streetscape The space between the buildings on either side of a street that defines its character. The elements of a streetscape include building frontage/façade, landscaping, sidewalks, street paving, street furniture, signs, awnings, and street lighting. Traditional neighborhood design (TND) A basic unit of new urbanism containing a center that includes a public space and commercial enterprises; an identifiable edge, ideally a 5-min walk from the center; a mix of activities and variety of housing types; an interconnected network of streets, usually in a grid pattern; and prominently located civic buildings and open space that includes parks, plazas, and squares. Transit-oriented development (TOD) A form of development that emphasizes alternative forms of transportation other than the automobile – such as walking, cycling, and mass transit – as part of its design. Transitoriented development locates retail and office space around a transit stop. This activity center is located adjacent to a residential area with a variety of housing options, such as apartments, townhouses, duplexes, and single-family houses. It is similar to a traditional neighborhood development. Urban growth boundaries Boundary lines around urban developments beyond which land development is not permitted. Zoning A legislative process that regulates building dimensions, density, design, placement, and use within different areas and districts of a community.

Sustainable Urbanism

Definition of the Subject Sustainable urbanism is walkable and transitserved urbanism integrated with highperformance buildings and high-performance infrastructure. Sustainable urbanism emphasizes that the personal appeal and societal benefits of neighborhood living – meeting daily needs on foot – are greatest in neighborhoods that integrate compactness (density), completeness, connectedness, and biophilia (human access to nature). Sustainable urbanism is really a call for integrating all human and natural systems that make up a neighborhood or corridor (groups of neighborhoods). It compiles many of the social, economic, and ecological systems that support complete neighborhoods and corridors. As with integrated building design, the magnified benefits may come at little or no additional cost. The locations with the greatest potential for cross-system integration are dense, mixed-use, and served by transit. Rights-of-way offer significant potential for better integration. Potentially immense economic and environmental benefits may result from integrating high-performance transport, water, sewer, lighting, and power systems with highperformance buildings that consume few to no resources and produce little to no waste. Sustainable urbanism is based on the neighborhood unit, an early planning construct. A performance-based definition of neighborhood is a settlement that has a defined center and an agreed-to extent, is walkable, and is diverse in terms of building types, people, and uses. Sustainable urbanism expands the role of the neighborhood to address its proportionate share of society’s social and environmental needs. Figure 1 shows a potential template for a sustainable urbanist neighborhood. This diagram builds on two previous neighborhood diagrams: Clarence Perry’s diagram of the neighborhood unit, published as part of the 1929 Regional Plan of New York and Its Environs [1], and the Duany Plater-Zyberk (DPZ) urban neighborhood diagram, which is based on Clarence Perry’s neighborhood unit. There are five distinctions in the resulting urban neighborhood diagram: (1) the sustainable neighborhood aspires to be a building block for a

Sustainable Urbanism

165

Sustainable Urbanism, Fig. 1 Sustainable neighborhood diagram. Doug Farr, Leslie Oberholtzer, and Christian Schaller (© Farr Associates)

transit corridor; (2) there is a fixed-route highintensity transit mode (BRT, trolley, light rail) at the center; (3) it is fitted out with high-

performance infrastructure, district power, dimmable streetlights, and a share car per block; (4) the mix and density support car-free housing

166

and a “third place” setting; and (5) habitat and infrastructure greenways offer the neighborhood an agreed-upon spatial extent. The importance of sustainable urbanism is pressing in the United States, where decades of single use zoning and sprawl have eliminated these historic distinctions. These distinctions are also urgent for emerging developments in Asia, Eastern Europe, Africa, and the Americas where US land use approaches are aspirational and considered signs of “progress.” For all of us, the guidelines and metrics of sustainable urban patterns will be critical.

Introduction Humans are now a “superspecies”; humans make personal and national choices that together determine the world that future generations will inherit, as well as the fate of many of the world’s other species. The American lifestyle puts society and the planet on the wrong course. Sustainable urbanism celebrates the power of design and offers an emerging pattern of human settlement that can strengthen the interdependence, and therefore the success, of all life on Earth. To understand how sustainable urbanism offers a solution, it is important to first understand the current problem. Americans are experiencing a serious deterioration in public health. The US population is sedentary and deprived of exercise, and the result is a rising incidence of obesity. Today’s average American weighs 25 pounds more than a generation ago [2]. While four-, five-, and even sixstory residential walk-up buildings were commonplace in most large American cities during the nineteenth century, the use of stairs has been actively discouraged by elevators and the fire stair enclosure requirements of twentieth- and twentyfirst-century US building codes. Lives are also increasingly lived indoors. New streetscapes are hostile to pedestrians and discourage travel by foot. New buildings are designed with air-conditioning for indoor living rather than operating with open windows and doors that draw people outdoors. The lack of human contact with nature has inured and possibly blinded people to the terrible

Sustainable Urbanism

damage being done to the planet. An unequivocal international scientific consensus backs the fact that, after only a few generations of the petroleum age, the resulting increase in human population and the increasing per capita impact from human activities have changed the Earth’s climate [3]. A prime reason for this, and a lifestyle choice made early and rarely questioned, is the love affair with the automobile. Most Americans rely on cars to meet their daily needs. The “freedom of the road” is cherished and safeguarded with a zealousness that suggests it was written into the Constitution. The joint addiction to driving and oil comes at an extremely high cost to individuals and families. As of 2016, the average cost of owning, operating, and maintaining a new car is estimated to be over $8,500 per year [4]. Parking exacts its own toll on business, government, and the environment. Street networks and parking spaces are expensive to build. The cost of constructing parking spaces is high ($2,500 to $5,000 for a surface spot and $30,000 to $50,000 for underground spaces) – a national capital investment of between $5 trillion and $10 trillion. Despite this enormous investment in parking, it is generally offered free to users, paid for by the private sector through increased business prices and by the public sector in taxes. Less than 1% of the 8,271,117 lane-miles [5] of highways, roads, and streets in the United States charge tolls [6]. Gas taxes pay most of the cost of highway construction and maintenance, while the vast majority of local roads are paid for with local taxes [7]. Furthering the challenge, from 1950 to 1990 Americans developed land at more than three times the rate of population growth [8]. This low-density development results in the highest per capita demands on natural systems and habitats. The conventional view in America is to think of cities as the source of the pollution that is causing climate change. Indeed, per unit of land area, cities generate a great deal of pollution (Fig. 2, traditional view). However, on a per capita basis, city dwellers generate the least CO2 (Fig. 2, emerging view). The American dream of a large house on a large lot in the suburbs is what’s most responsible for cooking the planet.

Sustainable Urbanism

167

Sustainable Urbanism, Fig. 2 Cities and CO2 (© Center for Neighborhood Technology 2007)

While energy codes adopted by states and municipalities over the last few years have increased building energy efficiency per square foot, the square footage of the average American house appears to be increasing more quickly, canceling out efficiency gains.

The public infrastructure required to support this low-density development is also expensive to build and maintain, especially when measured per capita. National studies show that low-density development increases the cost of hard infrastructure and with it the tax burden for all, by an average of 11% [9].

168

Despite the many benefits urbanism bestows on the Earth, conventional urbanism obliterates virtually all the systems of nature it comes into contact with. While the lower densities of conventional postwar suburbs allowed a higher percentage of the land area to consist of vegetation, much of it is residual and fenced into small private parcels, devaluing the utility of this unbuilt land for both human delight and nonhuman habitat. As a consequence of this suppression of nature, most people live out of daily contact with natural systems. Without feedback regarding the enormous stress that their lifestyle places on nature, they conduct their daily lives largely unconstrained by concerns about it. This disconnect from nature is increasingly thought to contribute to a number of psychological harms, such as increased stress and attention deficit hyperactivity disorder (ADHD) [10]. It should be clear now that the lifestyle choices made by Americans (and subsequently others around the world) have inexorably altered the built and natural environment. Americans are paying a terribly high price in individual health, a general sense of well-being, and happiness. People in the United States are alienated from sustenance-providing nature. Perhaps worst of all, the United States is jeopardizing the global climate and is confused as to the causes. Within the context of these wrongs, the goal of sustainable urbanism is to chart a compelling future.

Topic 1: Building a Movement Pioneering Reforms: Setting the Stage for Sustainable Urbanism Sustainable urbanism draws attention to the enormous opportunity to redesign the built environment

Sustainable Urbanism

in a manner that supports a higher quality of life and promotes a healthy and sustainable American lifestyle. The basis for this transformation of the built environment is a synthesis of urbanism – the millennia-old tradition of human settlements – with the late twentieth-century environmentalism that started with Rachel Carson’s Silent Spring. The synthesis of these two intellectual and practical histories requires a new consensus on the role of humans in nature. The 1969 book Design with Nature [11], by Scottish landscape architect Ian McHarg, was the first to explain to a relatively wide audience geographic information systems (GIS), the natural transect (Fig. 3), and other ecological principles. However, Design with Nature stopped short of trying to improve cities by fully integrating their design with natural systems. This obliviousness is revealed in the built work – essentially welllandscaped, auto-dependent suburbs – which are still mistakenly seen by many as a sustainable practice. Transcending McHarg, sustainable urbanism grows out of three late twentieth-century reform movements that highlight the benefits of integrating human and natural systems: the smart growth, new urbanism, and green building movements. Even these movements, highly worthy both individually and collectively, have a myopia when it comes to searching for long-term solutions. Further, there has been an understandable but unfortunate tendency toward self-validation, resulting in an unwillingness to engage in a larger, more comprehensive agenda. For instance, a certified green building isn’t really a positive contribution to the environment when it is surrounded by a massive paved parking lot, since a walkable neighborhood is hard to sustain when its houses are isolated.

Sustainable Urbanism, Fig. 3 The natural transect. Drawing by the office of Wallace, Roberts, McHarg, and Todd, circa 1970 (© Wallace Roberts, McHarg & Todd, LLC)

Sustainable Urbanism

Sustainable urbanism attempts to bring these three important movements together and knit them into a more comprehensive design philosophy to allow and create truly sustainable human environments.

Smart Growth: The Environmental Conscience of Sustainable Urbanism Smart growth has its roots in the environmental movement of the 1970s. The unique burst of federal environmentalism from that decade included the National Land Use Policy Act, which was intended to encourage states to develop coordinated state land use plans, and proposed a new federal agency and a land-planning database [12]. While the proposed act failed, its proposal for state-by-state land use planning was adopted in several pioneering states in the intervening years. In 1973 Oregon’s legislature passed a law requiring all the state’s municipalities to designate urban growth boundaries (UGBs), rings beyond which land development was not permitted [13]. UGB succeeded in controlling the scope of land development, thus preserving the state’s scenic treasures; however, it did little to ensure the quality of development within the UGB, leading to well-located bad development (aka “smart sprawl”). A piece of Maryland legislation, the Smart Growth and Neighborhood Conservation Program, was enacted in 1997 and designated urban growth areas that were eligible for state infrastructure. The smart growth movement embraced a broader agenda in 1996 with the development of ten principles of smart growth. At the time, many environmentalists were simply anti-growth and viewed all development, largely without distinction, as hostile to the environment. The principles were successful in uniting a decentralized grassroots movement of local and regional citizen activists and municipal leaders under the smart growth banner. However, the vagueness of the standards, and the smart growth movement’s decision to lend its name to development projects of sometimes minimal incremental improvement, worked to devalue the smart growth “brand.”

169

Congress for the New Urbanism: Sustainability’s Urban Design Movement The Congress for the New Urbanism (CNU) was founded by six architects in 1993 around a shared vision of promoting traditional urbanism as an antidote to conventional sprawl. Throughout the 1990s, new urbanism became an increasingly large part of mainstream development discourse. It has excelled at creating mixed-use neighborhood developments and transit villages, featuring town centers, fine-grained walkable street grids, and a highly diverse ensemble of traditional buildings and regional architectural styles. Because the projects are routinely deemed illegal under local zoning laws and go against most conventional development practices, the new urbanists have pioneered new approval techniques (notably the town planning charrette) and zoning recommendations [14]. The CNU desire to control the longterm placement and design of buildings led to the development of form-based coding, a highperformance alternative to conventional Euclidean zoning (Fig. 4). Despite its many achievements, the CNU has proved only somewhat successful in reforming state or national practices. In large part this is because the CNU has focused on convincing local regulators to create exceptions to conventional practice and to allow the approval of individual projects. While effective on a case-by-case basis, this pragmatic approach has left intact a foundation of hostile single-issue standards as well as a built environment that remains dominated by climate-changing sprawl. USGBC: Sustainability’s Building Performance and Certification Movement The oil shocks of the 1970s jump-started a movement for building energy efficiency and solarheated and solar-powered buildings. Inspired by the 1992 Rio Earth Summit, the American Institute of Architect’s Committee on the Environment published The Environmental Resource Guide in 1993, a comprehensive catalogue on the theory, practice, and technology of “environmental” buildings. This subsequently inspired the creation of the US Green Building Council (USGBC). The USGBC made two very smart moves to accelerate

170

Sustainable Urbanism

Sustainable Urbanism, Fig. 4 Three forms of development regulation: conventional zoning, design guidelines, and form-based codes (© Peter Katz and Steve Price)

Sustainable Urbanism

the adoption of environmental or green building practices: it expanded its audience outside the architecture profession, and it sought to mobilize the private sector. The USGBC drafted the pioneering Leadership in Energy and Environmental Design (LEED) standards for green building. The LEED standard combines prerequisites with optional credits that earn points toward an overall score, which then corresponds to LEED certification categories from Certified to Platinum. This flexibility works well in the marketplace, allowing a project to incorporate those green building strategies most attainable by the project. The US General Services Administration’s requirement for all government-owned and government-developed buildings to be LEEDcertified created a market for LEED-rated buildings and continues to deliver large square footages of LEED-certified projects every year. As a result, LEED has become increasingly mainstream, refocusing the entire building industry toward more sustainable practices. As of May 2016, there were about 80,000 projects participating in LEED and over 200,000 professionals with LEED credentials [15]. The backbone of the success of LEED has been the ability of the USGBC to quickly scale up – increasing its staff, its technical advisory groups, and certification operations at a geometric pace while maintaining quality and integrity. A second engine driving green building practice is the concept of integrated design: working in interdisciplinary teams to optimize overall building performance without adding construction cost. The dominant unit of reform within the LEED system remains the stand-alone building. This greatly limits the power of LEED certifications for individual buildings to have any effect on their surrounding context. In 2005, in a significant signal of its intention to move beyond the standalone building, the USGBC board modified its mission to address both buildings and community through the creation of LEED for Neighborhood Development (LEED-ND). The professional cultures of urban designers and architects are shifting; it is no longer acceptable to build a high-performance building in a

171

greenfield, automobile-dependent context and have it certified as “green.” It is also no longer good enough to develop in a responsible location and build an admirable, walkable, mixed-use neighborhood while ignoring the level of resources required to build and maintain the buildings within that location. The work and principles of the aforementioned groups and movements – smart growth, new urbanism, and green communities as represented by LEED-ND – are heartening developments. They are essential stepping-stones. Individually, however, none can solve the challenges that society is facing. Only with a concerted effort, only by fusing their various initiatives into a cooperative whole, can there be a new framework that supports a truly sustainable lifestyle. Pilgrimage Sites: Case Studies in Sustainable Urbanism The best sustainable urbanist neighborhoods are more than assemblages of energy-saving technologies. Sustainability embraces context. While cookie-cutter subdivisions and nondescript strip malls blanket the country like so many interchangeable parts, the best sustainable urbanism locks in to celebrate place. Sustainable urbanism takes stock of an area’s environmental, social, and economic wealth and augments it through neighborhood and corridor design. Case studies give the opportunity to see how various participants worldwide have adapted the principles of sustainability to the places they know and love. Developing sustainable urbanism requires a level of design coordination and of development sophistication beyond conventional practice. Yet pioneers have chosen to pursue this approach in projects worldwide. Amazingly, many projects needed only one well-placed champion to steer them in pursuit of a sustainable urbanist vision. These leaders can be mayors, planners, developers, activists, or any combination thereof. When so many barriers to the sustainable urbanist movement exist, vision and leadership are without doubt the most valuable assets these projects possess. One exemplary case study is the redevelopment of Uptown Normal. Since the year 2000,

172

Sustainable Urbanism

the Town of Normal, Illinois, has been implementing their transformative Uptown Renewal Plan. One component, Uptown Circle, became a community-defining plaza and green infrastructure innovation, treating Uptown’s stormwater sustainably. Uptown also kept up with changing sustainability benchmarks – first certifying as LEED-ND Silver in 2009, followed by becoming a Living Community Challenge Pilot in 2014. The project experienced so much success; the Town completed an Uptown 2.0 Plan in 2015 to explore strategies that would expand Uptown’s benefits to adjacent areas. When projects incorporate urban design, nature, and technology, they can become exemplary and can serve as powerful catalysts for continued improvements across the nation and globe. Part of the sustainable urbanist movement is maintaining a database of such exemplary projects. These can be found at www.thepatternproject.org. Lasting Impact The sustainable urbanism movement seeks to create a brand, agenda, and standards that can leverage further progress in the field of sustainable built environments. While no formal measures of impact exist, sustainable urbanism is moving the

field forward. For example, the 2012 book Sustainable Urbanism and Beyond continues the conversation about safe, healthy, and lovable urban environments and includes critiques of the sustainable neighborhood diagram (Fig. 1). The movement aims to accelerate the parallel reforms and efforts needed to create a tipping point in support of widespread adoption of sustainable urbanism. The progression from LEED certification for buildings to LEED certification for communities – a fundamental aspect of sustainable urbanism – has resulted in myriad new systems for evaluating and improving sustainable communities, including the Living Community Challenge, EarthCraft, STAR (Sustainability Tools for Assessing and Rating) Communities, EcoDistricts, etc. While there are plenty of projects that have the more generic “sustainable” label without actually producing sustainable outcomes, projects that meet the requirements of these certifications generally align with sustainable urbanist principles. Implementation: Steps and Leadership Agendas See Tables 1 and 2.

Sustainable Urbanism, Table 1 The three steps of sustainable urbanism The three steps of sustainable urbanism 1 agreeing to weights and measures: Making a market for sustainable urbanism

Citizen/resident/voter Having a standard for a certified walkable, healthy neighborhood adds value, like a seal of approval

Planner/developer I can use LEED-ND to build, develop, and market the best possible projects for my customers

2 dismantling petroleum-era barriers to sustainable urbanism 3 A national campaign to implement sustainable urbanism

Subsidies and zoning laws are benefiting special interests and are not in the best interest of my community Getting involved in a national campaign such as the 2030 community challenge makes me feel like I am contributing to something bigger

Density is no longer a dirty word but in fact provides opportunities for mass transit that can liberate people from their cars The 2030 community challenge is good for business and good for the planet

Elected official Creating incentives such as expedited permitting or reducing impact fees can encourage more sustainable developments that embrace rating systems such as LEED-ND Current codes are outdated and isolate America’s communities, forcing people to drive The 2030 community challenge can conserve scarce public money and help the environment. It’s just good government

Sustainable Urbanism

173

Sustainable Urbanism, Table 2 Implementation agendas for leaders Catalyzing action Promote sustainable urbanism

Hire sustainable urbanist professionals Select and support sustainable urbanist developers Benchmark sustainability goals

Revise outdated regulations

Initiate a catalytic sustainable urbanist project

Develop a sustainable neighborhood

Develop a sustainable corridor

Create places conducive to all forms of life

Implementation agenda Communicate the many benefits of sustainable urbanism. Update progress on sustainability Sponsor a sustainable urbanist lecture series to increase awareness and build the capacity to integrate sustainable urbanism into each and every project Hire staff and a development team with a sustainable urbanist outlook Require that an experienced sustainable urbanist be included on each project team Include sustainable urbanist developers on project teams Evaluate the level of public support for a given project Develop benchmarks for a municipality’s sustainability goals. Track progress toward those benchmarks regularly Modify comprehensive plan law to require sustainable initiatives and benchmarking of all sustainability goals in future municipal land use plans Replace parking, lighting, and building setback minimums with maximums Replace building height, density, and share car maximums with minimums. Adopt standards that address multiple variables on a per capita or sliding scale Find a model project of the appropriate size to emulate in your town Initiate a carshare program. Municipal planners should perform GIS analysis of neighborhoods to determine the number that can viably support car share. Increase this area by 5% per year Develop a car-free housing project Assess neighborhood completeness Municipal planners should map neighborhoods in terms of auto-dependent zones, pedestrian zones, and mixed zones. Aim to increase pedestrian zones by 5% per year Prepare a master land use plan identifying transportation, infrastructure, and wildlife corridors and the adjacent areas of influence Transit agencies should adopt “corridor transit warrants” that require the provision of public transit service when corridor land development conforms to minimum densities Meet demand for transit-oriented development by planning for regional corridors Integrate wildlife corridors into the regional network and protect them from development

Topic 2: Thresholds of Sustainable Urbanism Over the last 25 years, leading planners and urban designers have become sophisticated in their ability to conceive and sell attractive infill and masterplanned urbanist developments. The projects typically include networks of narrow streets, mixes of housing and other building types, and a variety of walk-to parks and address automobile parking in creative ways. These leading projects reveal thresholds, or rules of thumb, for designing and

developing sustainable urbanism. These dimensional or relational metrics, illustrated in the following sections, are based on expert judgment of what will satisfy “the 80% rule” guidance that applies most of the time and in most conditions.

Neighborhood Compactness: Increasing Sustainable Effectiveness through Density Many self-declared “neighborhoods” are either too small to support any land use variety or too large to be considered walkable.

174

Victor Dover addresses this problem by proposing a minimum neighborhood size threshold of 40 acres and a maximum of 200 acres, with a neighborhood center comprising between 6% and 10% of the total land (Table 3). Compactness, as compared to conventional development, is essential to achieving sustainable urbanism’s potential benefits for communities, regions, and the planet. Sustainable urbanism is simply not achievable at low densities, below an average of seven or eight dwelling units per acre [16]. Christopher Leinberger of the Brookings Institution echoes this: FARs (floor area ratios are the percentage of land covered by the equivalent of one story buildings) between 0.05 and 0.30, typical of current drivable suburban development, “do not efficiently support transit” and result in neighborhoods where “there are generally no destinations that are walkable on a day-in, day-out basis” [17]. For these reasons, sustainable urbanism requires minimum development densities of approximately eight dwelling units per acre (roughly four times higher than the average new US development density of two dwelling units per acre). Sustainable urbanism requires that any parking be sold separately from the dwelling unit and that a minimum of one carsharing vehicle and space be available for every ten dwellings (Table 4).

Neighborhood Completeness: Daily and Lifelong Utility For neighborhoods to meet both one’s short-term daily needs and one’s long-term needs over a lifetime, as well as to support robust life choices, they need to include a wide variety of land uses, building types, and dwelling types. Design to meet daily needs on foot creates universal independence at opposite ends of the age spectrum – the youngster who is not yet licensed to drive and the senior who can no longer drive – as well as for people with mobility impairments. Walkability and mixed-use are a critical aspect of completeness.

Sustainable Urbanism

Completeness also refers to the diversity of dwelling types needed to accommodate the varied needs for housing over a lifetime. Maintaining lifelong relationships with family and friends has been shown to increase health and longevity [18]. Aging in place allows relationships to be foot-powered, avoiding the expense and energy use of cars or airplanes. A neighborhood that provides a full range of housing types allows people and families to remain in the neighborhood even as their housing needs change. Research from Zimmerman/Volk Associates, Inc. shows that the optimum residential mix in a downtown or intown neighborhood (new construction or adaptive reuse) is 82% attached and 18% detached (Table 5). Traditional neighborhood design (TND) goes beyond walkability, density, and housing type to recommend: a center that includes a public space and commercial enterprises; an identifiable edge, ideally a 5-min walk from the center; a mix of activities and variety of housing types; an interconnected network of streets, usually in a grid pattern; and prominently located civic buildings and open space that includes parks, plazas, and squares. One of the primary advantages of traditional neighborhood-designed (TND) communities over conventional suburbs is the opportunity to walk to shopping and entertainment venues. However, few communities based on new urbanism have successfully implemented retail centers. Given the number of housing units common in America’s TND communities, retail centers fail to meet minimal sales necessary for its business owners to earn a reasonable income. Approximately 1,000 households are necessary to support the average corner store; convenience centers need about 2,000 households; and neighborhood centers require 6,000–8,000 households (Table 6). Neighborhood Connectedness: Integrating Transportation and Land Use Sustainable urbanism means that people have abundant opportunities to walk, ride, bike, and even use a wheelchair around the neighborhood, as well as having access to good transit service to adjacent neighborhoods and regional destinations.

80

148

142 89

Savannah, GA

Seaside, FL

Boston, MA

Queens, NY Palm Beach county, FL Maui, HI Flagstaff, AZ

108 151 40–200

50

Location Charleston, SC

6.4% 9.5%

2.8% 3%

7%

5%

9%

(6.9 acres) (14.4 acres) 3–10%

(4.1 acres) (2.7 acres)

(10.3 acres)

(4.1 acres)

(4.5 acres)

% Area devoted to center (acres) 9% (88 acres)

11 20

Unknownc Unknownc 350g 101g 342g

6,600b 800b 460 438 1,739 400 minh

7.2 9.96

Unknownc

330b 82.6

8.2

9.1

Unknownc

320b

Net residential density (DU/acre)a 7.6

Number of accessory dwellings Unknownc

Number of primary dwellings 5,428b

62,768 116,200

7,500e 18,000

708,319f

153,034e

180,200e

S.F. of commercial space Unknownc

1,586 1,417 100–400

52 390

4,785

1,912

3,604

Net commercial area (S.F./acre) Unknownc

b

a

Residential Density = Dwelling Units – Acres (which do not include roads, parks, and public areas); Source: ArcGIS 9 ESRI Data and Maps Source: US Census, 2000 c Data unavailable d The four wards chosen are bounded to the north by Rt. 25, to the east by Lincoln Street, to the west by Whitaker Street, and to the south by Oglethorpe Avenue e Source: ArcGIS Business Analyst f Source: Boston Redevelopment Authority g Potential number of units once completely built h This minimum is based on 2007 assumptions about conventional retail channels of distribution and per-household purchasing. The number will change as sustainable urbanism advances

Pulelehua Juniper point Optimum range

Name Historic city of Charleston Four wards in historic Savannahd Seaside (original 80 acres) The Boston north end neighborhood Forest Hills gardens Callery-judge grove

Size (acres) 1,015

Sustainable Urbanism, Table 3 Neighborhood definition

Sustainable Urbanism 175

176

Sustainable Urbanism

Sustainable Urbanism, Table 4 Sustainable urbanist thresholds for residential parking regulations. Policy Off-street parking spaces Reduced parking requirements For-sale parking spaces Shared-car parking and car On-street parking in front of development

Conventional practice Minimum number required per dwelling Not permitted

Sustainable urbanism Maximum number allowed per dwelling Provision for shared car with shared car replaces up to five off-street parking spaces Sold separately Minimum one per every ten dwellings Can be used to meet off-street or shared car requirements

Sold with dwelling None required Cannot be used to meet requirements

Sustainable Urbanism, Table 5 Optimum residential mix by housing type for downtown and intown neighborhoods (new construction and/or adaptive reuse)

Percent of all units

Range

Rental lofts/ For-sale lofts/ For-sale rowhouses/ apartments (%) apartments (%) townhouses/duplexes (%) 23–55 17–36 15–30

For-sale urban detached houses (%) 10–32

Average

37

18

25

To achieve internal connectedness, the entire neighborhood needs sidewalks on both sides of the street, and the distance between intersections needs to be relatively short, ideally no longer than 300–500 ft. The highest quotient of walkability will result when the buildings that shape the street space are set close enough to the front property line to spatially define the streets as public spaces, with a minimum degree of enclosure formed by a building-height-to-street-width proportion of 1:3 or closer. The majority of the street network should be designed for a maximum automobile speed of 20–25 miles per hour, and the widest street should have no more than two travel lanes between curbs (Table 7). Transportation demand management comprises physical aspects of the built environment, as well as behavioral aspects of how people conduct themselves within them. Physical measures of the built environment determine up to 90% of the trips generated in residential sites and up to 35% in nonresidential sites; behavioral measures determine up to 8% and 32% of trips generated in

20

100%

residential and nonresidential sites, respectively (Table 8). Parking price elasticities will vary according to context but generally range from 0.1 to 0.3, that is, every 10% increase in parking price results in a 1–3% decrease in parking demand. Biophilia: Connecting Humans to Nature Sustainable urbanism seeks to connect people to nature and natural systems, even in dense urban environments. People are three times more likely to walk along landscaped pedestrian routes [19]. Mature tree cover can further encourage daily outdoor activity by cooling outdoor summer temperatures between 5 F and 10 F [20]. It can also increase the value of adjacent real estate by 3–6% [20]. Regular walking can reverse agerelated brain deterioration [21]. Dense vegetation provides viable habitat for songbirds, adding an aural benefit. Taken together, the human benefits of living proximate to vegetation and habitats are immense. Among the most neglected realms in town planning are walk-to neighborhood parks and plazas. Parks and plazas with a high degree of

6–8,000

2,000

Dwellings necessary to support retaila 1,000

6 to 8

2

TNDS necessary to support retail (6 DU/Gross acre) 1

245

225

Sales per S.F. ($) 210

Highly varied from $7.25 to $40.00

$12–18

Average annual rent per S.F $14–16

1–2mile radius

1-mile radius

Average trade area Neighborhood (5-min walk)

4.0 cars/ 1,000 S.F of gross building area 4.0 cars/ 1,000 S.F. Of gross building area

Parkingb On-street

Mixeduse main street

Urban form Mixeduse corner building Main street

This number can be reduced significantly if the store is located along a major road with 15,000 cars per day and reduced nearly to zero if gasoline is sold Combines on-street and off-street parking

60,000–80,000

Neighborhood center

b

a

10,000–30,000

Convenience centers

Corner store

Gross retail area (S.F.) 1,500–3,000

Sustainable Urbanism, Table 6 Neighborhood requirements for various retail types

Supermarket, pharmacy, and video store

Specialty food market or pharmacy

Anchor stores Any smallscale retail

Sustainable Urbanism 177

a

4

2

Avenue

Street

25

25–30

Target operating speed (mph) 30–35

On-street parking lanes are 7–8 ft in width

Maximum through-traffic lanes 6

Street types Boulevard

10–11

10–11

Travel lane widthsa (ft) 11–12

Local routes

Transit Express and local routes Local routes Bike lanes

Bicycle facilities Parallel paths or bike lanes Bike lanes

Sustainable Urbanism, Table 7 Street types appropriate for low-speed urban contexts

Freight Regional truck routes Local truck routes Local deliveries No

Optional

Median Yes

Yes

Yes

Curb parking Optional

Yes

Yes

Driveway access Limited

Sidewalk

Sidewalk

Pedestrian Sidewalk

300–660

300–660

Intersection spacing (ft) 660–1,320

178 Sustainable Urbanism

Sustainable Urbanism

179

Sustainable Urbanism, Table 8 Site-level vehicle trip generation Residentiala

Nonresidential

Up to 55% Up to 9% 2% Up to 15% Up to 9% Up to 90%

N/A Up to 9% 2% Up to 15% Up to 9% Up to 35%

N/A No limit Up to 25% 25% reduction for transit service

Telecommutingc Other TDM programs

Up to 4% N/A N/A 25% reduction for transit service N/A N/A

Demand management subtotald

Up to 8%

Physical measures Net residential density Mix of uses Local-serving retail Transit service Pedestrian/bicycle friendliness Physical measures subtotal Demand management and similar measures Affordable housing Parking supplyb Parking pricing/cash out Free transit passes

No limit Up to 2%, plus 10% of the credit for transit and ped/bike friendliness Up to 32%

a

For residential uses, the percentage reductions shown apply to the ITE average trip generation rate for single-family detached housing. For other residential land use types, some level of these mitigation measures is implicit in ITE average trip generation rates, and the percentage reduction will be lower b Only if greater than sum of other trip reduction measures c Not additive with other trip reduction measures d Excluding credits for parking supply and telecommuting, which have no limit

Sustainable Urbanism, Table 9 Park proximity sales premium Distance to park in feet 100 300 600 1,300

Distance to park in miles (Approx.) 1/50 1/16 1/8 1/4

Round-trip walk in min 1 2.5 5 10

Sales premium 24% 15% 5% Insignificant

Source: Miller, Andrew Ross, “Valuing Open Space: Land Economics and Neighborhood Parks” MIT Department of Architecture Thesis, 2001

landscaping, a naturalized stormwater feature, or a view of the night sky play a key role in supporting biophilia. When designing parks or high-quality open spaces, consider the following standards: be within a 3-min walk of every dwelling, have a minimum park area of 1/6 acre, be bounded on at least two sides by public rights-of-way, and ideally not fenced and locked at night (only if necessary for security).

Parks also increase the price homebuyers are willing to pay to live close by, providing a very good return on investment. One study documented a willingness of homebuyers to pay up to a 24% premium for a house and lot facing a park or natural area (Table 9) [22]. In contrast to traditional stormwater engineering practices, which are designed to treat water as a waste product and direct water away from where it falls, sustainable approaches to site and regional

180

water resource management strive to consider water as a resource. The sizing of stormwater management solutions depends on the type and area served (Table 10). To strengthen each human’s interdependence with natural systems, sustainable urbanism believes that human settlements need to be designed to make resource flows visible and experiential. For example, a wastewater system that extracts nutrients to grow food in one’s neighborhood creates an incentive not to dump toxic chemicals down the drain. The ability to see and experience where resources are produced, and where they go after they are used, promotes a human lifestyle better integrated with natural systems. It is important to remember that treating all site and building water as waste carries a large energy burden for processing and redistributing our potable water. At approximately 800,000 kWhr/yr. per 1 million gallons demanded, the amount of pollution generated by a coal-powered electrical generating facility equals the amount of water pollution removed. Above that threshold, more pollution is being generated by the treatment facility than is being removed (Fig. 5). Finally, sustainable urbanism is committed to the ongoing livelihood of nonhuman species located in habitats close to human settlements. While sustainable urbanism recognizes the harm caused by human encroachment on natural habitats, it also recognizes the greater benefit of providing immersive and continuous nature within a reasonable walking distance of human settlements. Equally important to human and nonhuman existence is managing light pollution. Lighting design in a sustainable urbanist neighborhood uses light where it is most useful – at potential vehicle/pedestrian conflict zones, to accent building façades, and to light wayfinding elements. Reduce energy consumption and the adverse impacts of light on the nighttime environment by allowing only low ambient light in general urban neighborhoods, achieved by lighting level maximums, adjustable/dimmable lighting fixtures, and integrated lighting infrastructure (Tables 11 and 12).

Sustainable Urbanism

High-Performance Buildings and Infrastructure with Integrated Design In addition to connecting people to elements of the natural world, sustainable urbanism incorporates high-performance infrastructure and integrated design to help reduce the impact that urban areas have on natural ecosystems. Sustainable urbanism aims to achieve per capita building energy efficiencies through both square foot energy efficiencies and per capita energy efficiencies, often engaging district systems where appropriate. Multifamily dwellings found in urban settings engage shared walls and floors to reduce energy use on a per square foot basis in all climates with mechanical conditioning (Figs. 6, and 7). Following deep energy conservation in buildings, sustainable urbanism often incorporates district energy heating, cooling, or combined heat and power (CHP) systems to maximize energy efficiency. While there is no universal standard for the configuration of a district energy system that will be applicable in all settings, there are minimum requirements and ranges to consider when investigating the economic and technical feasibility of a district energy system: – For district heating systems, a minimum of 4,000 heating degree-days in a year. – For district cooling systems, customers typically should consume more than 1,000 equivalent full-load hours (aka a 200-ton peakdemand building should consume 200,000 ton-hours over the course of a year). – The maximum distance between a production plant and the end of the distribution network for an economical steam line is 3–5 miles; the maximum distance for a hot-water line when thermal energy is derived from a municipal solid waste incinerator is 3 miles and 15 miles when derived from an electrical power plant. – Sites predominantly occupied by newer buildings with existing in-building boiler and chiller equipment will not prove to be economical for a district energy system, as owners of these buildings may not be inclined to connect to the system.

Soil

Horizontal surface

Bioretention swale

Naturalized detention

Permeable paving

Soil

Hardscape

Bioretention rain garden

Centralized detention basin

Stormwater facility

Soil

Sustainable urbanist Soil

Conventional

Approach

Paving designed to allow water to pass through surface using porous asphalt or concrete or using interlocking concrete permeable pavers. Water can be stored in opengraded stone beneath the surface to meet local detention requirements

Yard depression planted with perennial vegetation. Includes layer of organic and sand amended topsoil above a gravel drainage layer (where needed) Depressed parking lot or roadside islands planted with perennial vegetation. Includes layer of organic and sand amended topsoil above a gravel drainage layer (where needed) Detention basin naturalized with shallow side slopes and native wetland and prairie vegetation

Excavated basin designed to temporarily detain stormwater runoff to meet locally defined allowable release rate

Brief description

Sidewalks and driveways

Edge of paving

Yard and adjacent to bottom of downspouts

Streets and alleys

Parkways and medians

Area served by facility Lot Block

Sustainable Urbanism, Table 10 Conventional and sustainable urbanist stormwater management facilities

Stormwater parks, parkway medians

Neighborhood

(continued)

8–12% of site area. Size may be reduced where upstream bioretention, permeable paving, or greenroofs are utilized Net paved area

10–15% of impervious area – Less for permeable sandy soils

10–15% of roof area – Less for permeable sandy soils

8–12% of site area

Rule of thumb for sizing stormwater facilitya

Sustainable Urbanism 181

a

Stormwater facility

Extensive greenroof

Intensive greenroof

Horizontal surface

Roof

Roof

Vegetated roof with drought-tolerant species requiring little or no inputs for vegetative maintenance. Typically 3 to 4 inches of growing medium, depending on vegetation Vegetated roof with a wide range of vegetation, including grasses, shrubs, and even trees. May require irrigation and fertilization. Typically 8 inches and deeper growing medium, depending on vegetation

Brief description

Varies based on local stormwater standards and site imperviousness

Approach

Sustainable Urbanism, Table 10 (continued)

Building roof

Area served by facility Lot Block Building roof

Neighborhood

Net buildable roof area

Net buildable roof area

Rule of thumb for sizing stormwater facilitya

182 Sustainable Urbanism

Sustainable Urbanism

183

Sustainable Urbanism, Fig. 5 Typical energy consumption of various wastewater treatment systems (© 2007 Tom Ennis) Sustainable Urbanism, Table 11 Descriptions and general lighting allowances for lighting zones Transect zone Allowed initial lamp lumens/SF Base allowance (lumens) Lighting design criteria

LZ0 Rural and reserve 1.25–1.6a

LZ1 Reserve and suburban 2.5–3.2

LZ2 General neighborhood 3.3–4.2

LZ3 Urban center

LZ4 Urban core

7.6–9.7

10.9–13.9

0 No ambient light

17,000 Very low ambient light

24,000 Low ambient light

44,000 Medium ambient light

60,000 High ambient light

Chart data compiled from Model Lighting Ordinance (draft), Illuminating Engineering Society of North America (IESNA), and International Dark Skies Association (IDA) a This minimal lighting should be turned off most of the time

Sustainable Urbanism, Table 12 Current and ideal lighting practices Public expectation Light level regulations Control technology Control Roadway lighting Human links to nature

Current practice Outdoor brightness Minimums On-off Municipally controlled Pole-mounted Glare obscures all but a handful of stars

2030 sustainable urbanist ideals Outdoor darkness Maximums Addressable ballasts allow nighttime dimming Block and neighborhood controls Incorporated into roadway fabric Milky way visible across North America

184 Sustainable Urbanism, Fig. 6 Building massing and orientation types included in an energy model (© Alan Chalifoux)

Sustainable Urbanism

DETACHED, SINGLE-FAMILY

OPTION 1 S/V = 0.21

GLAZING = 25% WALL AREA



64

25



10’ FLOOR HEIGHT, TYPICAL ALL OPTIONS

OPTION 2 S/V = 0.21

64





25

OPTION 3 S/V = 0.20 40



40



TOWNHOUSE, ONE PARTY WALL

OPTION 4 S/V = 0.15

GLAZING = 30% WALL AREA



25

32



OPTION 5

25

TOWNHOUSE, TWO PARTY WALLS



32



S/V = 0.11

GLAZING = 35% WALL AREA

OPTION 7 ADDS SHADING

OPTION 6 S/V = 0.11

OPTION 8 ADDS SHADING

Future Directions The process of developing and implementing sustainable urbanism exposes areas of research necessary to advance the movement. For example, the impacts of sustainable urban design on humans are not well qualified. The field is in desperate

32





25

N

need of novel approaches to measuring the urbanism factors of greatest significance for human and environmental health outcomes. It is still unclear what impact local projects have on their surrounding communities, as well as whether and how they contribute to regional gains. Arguably most important is the question of whether sustainable

Sustainable Urbanism

185

Sustainable Urbanism, Fig. 7 Annual energy use for building types across US climate zones. Energy model results (© Alan Chalifoux)

urbanism projects are having their intended positive effects on both humans and the environment. Sustainable Nation: Urban Design Patterns for the Future (2018) offers more rules of thumb for how to achieve sustainable neighborhoods across the United States and beyond. While change must happen faster than ever before in order to avoid the lasting negative effects of climate change, there has never been a better time to unite around common passions and move forward on creating positive change neighborhood by neighborhood.

Bibliography Primary Literature 1. Thomas A, Delano FA (1929) Regional Plan of New York and Its Environs. The Graphic Regional Plan: Atlas and Description, (vol 1). http://babel.hathitrust.org/cgi/ pt?id=mdp.39015053244870;view=1up;seq=13 2. Centers for Disease Control and Prevention (2016) Obesity and overweight data for U.S. Adults. National Center for Health Statistics. Last Updated June 2016. http://www.cdc.gov/nchs/fastats/obesityoverweight.htm 3. Rosenthal E, Revkin AC (2007) Science panel says global warming is ‘Unequivocal’. New York Times 4. American Automobile Association (2016) Your driving costs 2016. http://publicaffairsresources.aaa.biz/ wp-content/uploads/2016/03/2016-YDC-Brochure. pdf. Accessed 19 June 2017, p 7

5. U.S. Department of Transportation, Federal Highway Administration, Conditions and Performance Report, Exhibit 2-7 (2002) http://www.fhwa.dot.gov/policy/ 2002cpr/Ch2b.htm 6. World Bank Group (2007) Toll roads and concessions. http://www.worldbank.org/transport/roads/ toll_rds.htm. Accessed 5 Feb 2007 7. Litman T (2007) Whose roads? Defining bicyclists’ and pedestrians’ right to use public roadways, Victoria Transport Policy Institute, November 30, 2004. http:// www.vtpi.org/whoserd.pdf. Accessed 5 Feb 2007, p 6 8. Author calculation, Wiley book. 9. Costs of Sprawl (2000) TCRP report 74, transit cooperative research program, transportation research board, national research council, 2002. http:// onlinepubs.trb.org/onlinepubs/tcrp/tcrp_rpt_74-a.pdf. Accessed 5 Feb 2007. Calculations done by author from data on pp. 222, 249 10. Louv R (2005) The last child in the woods: saving our children from nature-deficit disorder. Algonquin Books, Chapel Hill, pp 104–105 11. McHarg IL (1969) Design with nature. Wiley, New York, 1992 , p. v 12. Government Law Center, Albany Law School (2007) Smart growth and sustainable development: threads of a national land use policy, spring 2002. http://www.governmentlaw.org/files/VLR-Smart_gro wth.pdf. Accessed 13 Feb 2007, p 4 13. Oregon State Senate, Oregon Land Use Act (SB 100), enacted 1973 (2007) http://www.oregon.gov/LCD/ docs/bills/sb100.pdf. Accessed 5 Feb 2007 14. Lennertz B, Lutzenhiser A (2006) The charrette handbook: the essential guide for accelerated, collaborative community planning. American Planning Association, Chicago 15. Shutters C, Tufts R (2016) LEED by the numbers: 16 years of steady growth. http://www.usgbc.org/

186 articles/leed-numbers-16-years-steady-growth. Posted in LEED, May 27, 2016 16. LEED for Neighborhood Development requires a minimum of seven dwelling units per acre. U.S. Green Building Council, LEED for Neighborhood Development Rating System (Pilot Version), Neighborhood Pattern and Design Prerequisite 2: Compact Development. http://usgbc.org/ShowFile.aspx?DocumentID= 2310. Accessed 24 Feb 2007 17. Leinberger CB (2007) Back to the future: the need for patient equity in real estate, Brookings Institution, January 2007. http://www.brookings.edu/metro/pubs/ 200701226_patientequity.htm. Accessed 24 Feb 2007, p 3 18. Ewing H, Kreutzer R (2006) Understanding the relationship between public health and the built environment, https://www.usgbc.org/resources/understandingrelationship-between-public-health-and-built-environ ment-report-prepared-lee, p 92, U.S. Green Building Council, 2006.

Sustainable Urbanism 19. Simmons M, McLeod KB, Hight J “Healthy neighborhoods,” Chapter 7: Sustainable urbanism: urban design with nature, John Wiley & Sons, 2008 20. Center for Urban Horticulture, University of Washington, College of Forest Resources (1998) Urban forest values: economic benefits of trees in cities, November 1998. http://www.cfr.wash ington.edu/research.envmind/Policy/EconBens-FS3. pdf. Accessed 25 Feb 2007 21. News Bureau, University of Illinois at UrbanaChampaign (2006) Exercise shown to reverse brain deterioration brought on by aging, November 20, 2006. http://www.news.uiuc.edu/news/06/ 1120exercise.html. Accessed 25 Feb 2007 22. Miller AR (2007) Valuing open space: land economics and neighborhood parks, master’s thesis, Department of Architecture, Massachusetts Institute of Technology. http://dspace.mit.edu/handle/1721.1/8754. Accessed 26 Feb 2007

Water and Sustainable Design Herbert Dreiseitl DREISEITL consulting GmbH, Überlingen, Germany

Article Outline Glossary Introduction Water Challenges to Be Met by the Built Environment Water Contributions to the Built Environment Water Technologies and Functionality for a Sustainable Future Innovative Waterscapes and Blue-Green Infrastructures BGI Benefits for Health and Well-Being: BishanAng Mo Kio Park Conclusion: Water’s Position in Sustainable Urbanism Ten Guidelines for Working with Water in Sustainable Urban Design Bibliography

Glossary Blue-green infrastructures (BGI) new professional practices that integrate the diverse disciplines needed for comprehensive water design through an approach to water management that protects, restores, and emulates the natural water cycle with extensive landscape (green infrastructure) improvements. Deluge period of abnormally high rainfall that may result in severe flooding. Drought a prolonged period of abnormally low rainfall that may result in a shortage of water. Green infrastructures an approach to water management that protects, restores, and emulates the natural water cycle.

Technical water systems the integration of design and engineering of green and gray infrastructures, into solutions for deluge, drought, and water contamination. Technical water systems combine water retention and infiltration techniques with water storage, treatment, and reuse through the integration of green (natural landscape) and gray (constructed) systems. Water contamination water that is fouled by agricultural, industrial, construction, and human waste due to inadequate containment, aggravated by flooding. Water-sensitive regenerative cities the successful integration of water at the beginning of the planning process, comprehensively and beautifully designed. Watersheds the natural path of rainfall and stream flows given the topographic and geologic construct of the land, flowing into rivers, lakes, wetlands, or the sea.

Introduction Good water is even more urgently important to us than our daily food. Water in the built environment is everywhere [1], but not always visible and often not in our thoughts. In architecture and engineering, we manage and control water in an infrastructure of pipes, canals, and storage facilities that are mostly hidden and therefore not in the day-to-day consciousness of the average citizen. We take water’s services for granted, making use of its enormous qualities and performance attributes to produce and transport energy, regulate temperatures, enable chemical reactions including the production of concrete, transport waste and trash, produce food, and many other useful services. The traditional urban fabric systems that control, regulate, and enable water flows and metabolic processes are deeply interwoven and depend on a permanent and steady water supply; and the balance of not too much and not too little water is the focus of those technical infrastructures.

© Springer Science+Business Media, LLC, part of Springer Nature 2020 V. Loftness (ed.), Sustainable Built Environments, https://doi.org/10.1007/978-1-0716-0684-1_1032 Originally published in R. A. Meyers (ed.), Encyclopedia of Sustainability Science and Technology, © Springer Science+Business Media, LLC, part of Springer Nature 2020, https://doi.org/10.1007/978-1-4939-2493-6_1032-1

187

188

In contrast to these traditional urban fabric systems, water in the natural environment is never static but in a state of constant flux. In reaction to changing weather and precipitation, the water flow has a dynamic interaction with the terrain and greenery and is the environment’s resilient language with potential to create living systems and enhance evolution. The blue and green elements in nature are the drivers for a living landscape and a sustainable natural environment. The almost unpredictable, ever-changing characteristics and processes of water in the landscapes have always challenged the human desire for beauty, safety, comfort, and independence. Yet innovative responses in architecture are only just emerging. In today’s urban fabrics, buildings of any function are mostly conceived as shelters. They separate an inner space from the outer environment. For the first time in history, due to our very high standards of insulation, air control, integrated intelligent control systems, light regulation, etc., we can live in buildings in the urban fabric for weeks and months almost without any contact with the natural environment. Cities are like perfect machines with supporting infrastructures that seem to make the natural environment a by-product. Until now, dependency on nature has been treated as a relic of the past. Today, and even more so in the future, we can clearly see how interrelated humans are with the nature that surrounds them.

Water Challenges to Be Met by the Built Environment With the increasing intensity of hurricanes, superstorms, cloudbursts, and the rise in sea level countered by heat waves and droughts, the waterrelated environmental challenges are increasing for the built environment. In light of those environmental challenges, sustainable water concepts and systems that work interactively with the water regime are gaining in public awareness. Sustainable designers have begun to map the historic watersheds and landscape mix that supported locations for millennials. It seems that today we are at a turning point and

Water and Sustainable Design

more than ready for integrated water solutions in the sustainable built environment. If we compare infrastructures in the natural environment with those in urban settings, a significant difference appears in most cities today: Natural structures work with flexible space assignments and resilient principles. One of the characteristics of nature’s water management is the dynamic reaction and balance to a range of events, from a subtle change in water availability to an unexpected deluge or drought. Ultimately, it is all about allowing processes to develop over time and having enough multifunctional space, such as retention areas, bioswales, constructed wetlands, and others to operate as a dynamic buffer zone.

Water Contributions to the Built Environment One of the main characteristics of water systems is the ability to recover from climatic extremes, enabling biodiversity and ecosystem services. In the last few decades, nature’s responses to stress have become highly relevant and a focus for innovative architects, landscape architects, and engineers. Engaging blue-green infrastructures to recover from environmental stress while simultaneously providing better air, balanced acoustics, increased biodiversity, and finally livability is the subject of many pilot projects around the world. These projects foster awareness and encourage architects and other disciplines to overcome the traditional silos in the profession. Today, architecture for sustainable built environments demands a true paradigm shift for water design and is making huge critically needed steps forward. The design approaches and technologies for sustainable water management are being extensively explored and tested with success, to be further illustrated.

Water Technologies and Functionality for a Sustainable Future Wherever rainwater and stormwater touch the building or the ground, sustainable water

Water and Sustainable Design

management must begin, since all surfaces are a collecting facility, as well as a potential source of pollution. Therefore, the management of these areas, their design, and maintenance are of upmost importance. The materials selected for rooftops and surfaces influence the quantity and quality of water. The vegetation of green roofs, for example, can filter out pollution like heavy metals or particulates that are harmful to water supplies. Water retention and some level of evaporation and infiltration in the design of roof and ground surfaces can avoid the peaks of runoff flow and calm the erosive process. In addition, captured rain and stormwater are the best resources to recharge the groundwater aquifer and refill lakes and rivers and, therefore, should be filtered and released in the best quality. Rainwater should be collected, harvested, and reused [3] for irrigation at a minimum or used to subsidize the drinking water as in Singapore’s ABC Water Program (Figs. 1 and 2). In addition to the management of storms and rainwater, cities also need to be designed for climate resiliency to anticipate sea level rise. The CloudBurst program [4] of Copenhagen recommends design strategies to drain stormwater out to the sea, to store stormwater, and to make buildings resilient in the face of rising water. Since fresh water of drinkable quality is becoming a limited resource in many more regions of the world, we must carefully use and recycle water matched to each purpose. There is no reason to use processed, potable drinking water

Water and Sustainable Design, Fig. 1 Rainwater should be collected, harvested, and reused. Diagram of Stormwater ABC Water Guidelines. (Credit: Herbert Dreiseitl)

189

to transport our excrements and trash. Every drop of potable water should be used at least three times – for drinking, for gray water uses such as irrigation, and for black water uses such as toilets. Sustainable designers seek the complementary actions of avoiding the sources of water pollution and reclaiming the water, transport, energy, and nutrient qualities of each drop of water – the metabolic processes inherent in nature. Eighty percent of the water in the built environment is used for our food production with enormous energy consumption and contamination that leave a very negative footprint. Most of that negative footprint is not visible in the urban environment, resulting in almost no awareness by the final consumers. Sustainable solutions, those with a nutrient-holistic and water-sensitive approach, demand a different form of agriculture in both rural and urban areas. There are important examples of this approach including the “Urban Farmers” in Basel and the Netherlands, the “Aquaponic Gardens” in New York City and Singapore, and the “Biodynamic Farm movement” including the Hawthorne Valley in New York State. Water also has a critical social and quality of life function. Water catchment systems can be interactive and communicate the power and beauty of nature to citizens. To work sustainably with water, we need to create multifunctional and shared systems and spaces. Smart, flexible shared places such as plazas, parks, and fields can absorb and contain water after a big storm but allow recreation and social activities the next day.

190

Water and Sustainable Design

Water and Sustainable Design, Fig. 2 Both quality and quantity controls are key to rainwater collection, harvesting, and reuse. BGI Toolkit. (Credit: Giovanni Cossu (LCL Ramboll))

Scharnhauser Park in Stuttgart and Bishan-Ang Mo Kio Park in Singapore demonstrate this in a very convincing way. In 2006, the Public Utilities Board and National Water Agency of Singapore began a program entitled ABC Waters – Active, Beautiful, and Clean – that aims to realize the full potential of an integrated blue-green infrastructure (BGI) approach. By treating rainwater as a prime resource to fill reservoirs and water bodies, the ABC Waters’ program is a strategic initiative that works with the entire urban catchment of the island. Instead of channeling rainwater away from the city, areas within the urban development are used for rainwater collection. This also helps to contribute to the country’s water security. The program was implemented with the expectation that it will lead to multiple projects by private developers. Estimated 150 projects are on their way to be realized in the next 20 years. One of the pilot projects, and by far the largest, was the transformation of 3.2 km of the Kallang River and 62 hectares of Bishan-Ang Mo Kio Park (Fig. 3).

This is a classic BGI project [5] with a strong social component of over 3 million visitors per year. The design was carried out under the leadership of landscape architects in an interdisciplinary planning process, which included on-site test studies and hydraulic modeling with flow simulations. This BGI was designed to accommodate the dynamic process of a natural river system, which includes fluctuating water levels and widths to make sure unexpected problems are accounted for. A special focus on security led to the creation of a special safety system in case of sudden water rises in the open river valley. Elements from the concrete canal – previously on the site – were recycled and reused as substrate in the riverbed and on a specially formed platform for artwork. Today, the park comprises a vibrant urban river with natural elements, although it is still very much formed and shaped by the people that use it. Better hydrological capacities, upgrading parameters of limnology, efficient erosion control, and other advances were possible through appropriate bioengineering techniques and reshaping of the river profile (Figs. 4, 5, and 6).

Water and Sustainable Design

191

Water and Sustainable Design, Fig. 3 The 62 hectares of Bishan-Ang Mo Kio Park, Singapore, has been transformative for the environment and quality of life. (Photo Credit: Herbert Dreiseitl)

Water and Sustainable Design, Fig. 4 Fostering closer connection with water (Bishan-Ang Mo Kio Park Photo Credits: Herbert Dreiseitl)

Plants and bedding materials were used to stabilize the banks to withstand the erosive energy of high water flows while at the same time creating

diverse stream habitats for native plants and animals. Not only is there a significant increase in biodiversity, there is also a completely new

192

Water and Sustainable Design

Water and Sustainable Design, Fig. 5 Quality of life gains are matched by better hydrological capacities, erosion control, and other advances in water management and use. (Bishan-Ang Mo Kio Park Photo Credits: Herbert Dreiseitl)

atmosphere that has resulted in connecting the people in the neighborhood together. Socioeconomic factors play into this development, and a detailed study was done on the effects of the BGI on real estate values and other parameters. The Bishan-Ang Mo Kio Park has won numerous awards and is seen as a new vision and role model for BGI which addresses the dual needs of water supply and flood management while creating spaces for people and nature in the city.

Innovative Waterscapes and Blue-Green Infrastructures The emergence of new professional practices that integrate the diverse disciplines needed for comprehensive water design is contributing to the literature on innovative waterscapes. The following excerpt from Strengthening Blue-Green Infrastructure in our Cities [6] captures the philosophy and design potential of innovative waterscapes

proposed by Herbert Dreiseitl and Bettina Wanschura (Ramboll Liveable Cities Lab), Matthias Wörlen and Manfred Moldaschl (Zeppelin University) Nirmal Tulsidas Kishnani and Giovanni Cossu (NUS National University of Singapore) and James Wescoat and Karen Noiva (MIT). Blue-Green Infrastructures as Tools for the Management of Urban Development and the Effects of Climate Change The conventional approach to urban water infrastructure has been to use quantitative models to predict future water demand and then to construct additional infrastructure to meet this demand. This approach prioritizes technology and large physical interventions which attempt to manipulate natural processes to suit the needs of humankind. However, the focus on “gray” infrastructure – so-called because of the massive amounts of concrete and metal typically involved – is progressively showing deficiencies

Water and Sustainable Design

193

Water and Sustainable Design, Fig. 6 Symbotic relationship between Urban Infrastructure and Natural Systems. (Bishan-Ang Mo Kio Park Photo Credits: Herbert Dreiseitl)

and limitations in meeting the additional stresses to urban water supply and management induced by rapid urbanization, impervious land cover, and climate change. In some cases, the reliance on gray infrastructure can actually contribute to these stresses. For instance, the conventional approach to urban stormwater runoff has been to collect precipitation in a connected sewer system and to transport it out of the city as quickly as possible [7]. As cities have grown, impervious land cover has increased which generates a larger volume of stormwater runoff in a shorter period of time, overwhelming existing sewers and increasing flooding. Gray infrastructure will also fail to mobilize the many potential socioeconomic benefits of water in enhancing the aesthetics of the urban fabric and the quality of life. In response to these changing times, decisionmakers are starting to look beyond the gray and experimenting with less conventional approaches to infrastructure. Blue-green infrastructure (BGI)

offers a feasible, economical, and valuable option for urban regions facing challenges of climate change. It complements and in some cases mitigates the need for gray infrastructure. BGI represents a paradigm shift that recognizes the importance of and value in including the role of urban hydrology within urban water management. The “blue” recognizes the importance of the physicality of water itself, while the “green” connects urban hydrological functions with vegetation systems in urban landscape design. The resulting BGI has overall socioeconomic benefits that are greater than the sum of the individual components (Fig. 7). In this context, the Liveable Cities Lab (LCL) initiated a research project “Enhancing BlueGreen and Social Performance in High Density Urban Environments” [8]. The goal of this research was to move toward a more comprehensive understanding of underlying concepts contributing to the effective implementation of BGI. This entry summarizes the key results of the

194

Water and Sustainable Design

Water and Sustainable Design, Fig. 7 The critical shift from gray to blue-green infrastructures in Bishan-Ang Mo Kio Park, Singapore, in 2008 and 2013. (Photo Credit: Atelier Dreiseitl)

project and focuses on challenges, obstacles, and successes of selected BGI case studies. The Definition of BGI The topic of green infrastructure is now a wellestablished concept in urban environmental planning, policy, research, and design, and awareness and understanding of its potential benefits for ecology and society have increased. The term green infrastructure often refers to projects that include vegetated design elements such as parks, green roofs, greenbelts, alleys, vertical and horizontal gardens, and planters. Such green infrastructures are recognized and intensively discussed with respect to the ecosystem services they provide – services that are especially valuable in densely populated urban areas [9]. However, “green” infrastructure is a bit of a misnomer, as infrastructures of this type are often closely linked with and even defined by “blue” processes. Blue infrastructure technically refers to infrastructure related to the hydrological functions, including rainwater and urban stormwater systems as well as surface water and groundwater aquifers. In urban design blue infrastructure is traditionally discussed as a matter of resilient provision for water supply and water security. Such

water infrastructure may be natural, adapted, or man-made and provides functions of slowing down, decentralization and spreading, soaking into the underground, and evaporating and releasing water into the natural environment. This must critically include flow control, detention, retention, filtration, infiltration, and different forms of water treatment like reuse and recycling (Fig. 8). In general, blue infrastructure addresses aspects of water quantity as well as quality control. The BGI paradigm marries these two types of infrastructures and values together in a union that is greater than the sum of its parts. Benefits of BGI BGI integrates hydrological and biological water treatment flows into systems where green features are seamlessly overlapping with blue features. Together blue and green infrastructures strengthen urban ecosystems by evoking natural processes in man-made environments and combining the demands of sustainable water and stormwater management with the demands of urban planning and urban life. As a result, BGI systems have positive impacts on the urban metabolism of natural resources (added green values) and on the experience and behavior of people using these

Water and Sustainable Design

195

Water and Sustainable Design, Fig. 8 BGI must critically include flow control, detention, retention, filtration, infiltration, and different forms of water treatment.

Neighbourhood Scale Solutions. Giovanni Cossu (LCL Ramboll))

infrastructures (added social values). A selection of the benefits associated with the implementation of BGI in dense urban areas is presented below.

Quantity-related benefits of BGI include (i) BGI enhances on-site retention of stormwater, which protects valuable wetland areas, reduces the need for designated downstream areas for flood buffer zones, and reduces the risk and impact of flooding; (ii) the natural unsealed surface allows water to seep into the ground, recharging underlying aquifers and balancing the groundwater level.

Water-Related Benefits

BGI effectively controls the quantity of stormwater but also improves water quality. Quality-related benefits of BGI include the following: (i) plant roots in combination with soil absorb nutrients and purify infiltrating water and also improve the general water quality in urban catchment areas, thereby reducing energy demands and costs associated with water treatment; (ii) BGI contributes to the avoidance of overheating and oxygen shortage caused by high temperatures of concrete materials.

(Graphic

Credit:

Climate Change Adaptation and Biodiversity

Besides benefits directly related to water and plants, BGI has a huge potential to modulate the urban climate by reducing urban heat island effects, balancing diurnal temperature fluctuation, and supporting natural air ventilation.

196

BGI also reduces the bioclimatic impacts of land cover changes such as desiccation of urban soils and associated wind-borne air pollution and dust hazards. By managing and modulating hydroclimatic variability and weather extremes, BGI enhances the adaptability and resilience of urban infrastructures. BGI also increases urban biodiversity as it improves rich biotopes and landscape connectivity, protects aquatic ecosystems, and creates biodiversity-rich zones to sustain flora and fauna. BGI Enhances a City’s Beauty and Aesthetics

BGI helps to reconnect people with the natural environment through the active integration of water and greenery in which the boundaries between the two are blurred and made accessible. Blue elements of urban design tend to have the strongest positive associations, and when combined with green elements, this positive effect is magnified. The perception of the relative beauty of the blue elements seems to be related to their scale and size, as well as how the edge conditions for public access are implemented. Societal Benefits of BGI

BGI creates upgraded space for recreation, exercise, and social activities and therefore helps to improve human physical and mental health. These amenities reduce individual and public health costs. BGI supports social interaction and social integration as it increases the tendency to use open spaces for activities in groups and the commitment to spend time with families, neighbors, and communities. By improving social and aesthetic attractiveness of surrounding land and buildings, BGI increases property values and real estate values. The creation of blue-green infrastructure signals a city’s overall attractiveness and livability and increases the reputation of a city’s governmental institutions to take care of their residents’ living conditions. Finally, BGI supports biophilia – people’s affinity with nature – as it reconnects people with natural forms, elements, and processes that have major benefits for human happiness and the willingness to protect nature.

Water and Sustainable Design

Main Challenges for Successful Implementation of BGI in Dense Urban Areas The main constraints on implementing sustainable urban stormwater and environmental management in a changing climate are not technological. Rather, they involve shifts in vision, policy, design, and the urban planning culture. The transition of urban water management from standard gray to blue-green implies a change in the social and political setting of a city. BGI relies on the capabilities in a city to negotiate forms and outcomes with all civic stakeholders, as well as to be aware of unintended consequences in the wider context (spatial, social, temporal). As BGI is still a rather unknown approach in many cities, practitioners, politicians, and citizens have to be convinced that BGI is able to guarantee at least the same level of security as older established solutions and that it can provide new types of security for climate resilience. Water planners otherwise tend to fall back upon the gray infrastructure approaches followed under historical climatic conditions or install redundant blue and green infrastructure elements at low levels with higher costs to avoid risk. This has limited the wide implementation of BGI elements and techniques to achieve multifunctional urban landscapes on a holistic catchment scale. BGI is a valuable and viable opportunity for creating multifunctional landscapes with an ecological approach to sustainable urban stormwater practice (Fig. 9). A paradigm shift is needed to ensure that urban water management moves beyond the conventional engineering mind-set to a more holistic approach that includes knowledge about societal values and ecosystem services. Such a paradigm shift has begun to be appreciated, but many decision-makers still remain unaware of the value of BGI or how to operationalize it. BGI Case Studies In order to provide a more balanced picture of BGI challenges relevant around the world and in a variety of contexts, the Liveable Cities Lab used several selection criteria to generate in-depth case studies, including climate, governance systems, and variations in the history of BGI development

Water and Sustainable Design

197

Water and Sustainable Design, Fig. 9 BGI creates multifunctional landscapes with an ecological approach to sustainable urban stormwater practice. (Building-Scale Solutions Graphic Credit: Giovanni Cossu (LCL Ramboll))

types, as well as the designed functionality within the BGI. The cases chosen for the study represent several continents (America, Europe, and Asia) and a range of climate types including the tropical rainforest climate (Singapore), the tropical wet and dry climate (Mumbai), and the humid continental climate (Germany, Denmark, etc.). For each case study, positive and negative lessons were identified and an attempt made to generalize these lessons as good practices important for current and future BGI planning and implementation in cities.

Case studies on the project level included the following: Emerald Necklace, Boston (USA); Hannover-Kronsberg (Germany); Bishan-Ang Mo Kio Park (Singapore); Khoo Teck Puat Hospital and Yishun Pond (Singapore); and Ulu Pandan Park Connector (UPPC) (Singapore). Case studies on city level included Hamburg (Germany); Portland, Oregon (USA); Copenhagen (Denmark); New York City (USA); Jakarta (Indonesia); and Mumbai (India). A selection of these case studies is presented below.

198

BGI in Hannover-Kronsberg, Germany

Hannover-Kronsberg (Germany) is a residential area with 3000 dwellings built from 1992 to 2000 as an exhibit for the World Exposition 2000 titled “Mensch-Natur-Technik” (HumanNature-Technology). Following Agenda 21, the Habitat II Model, and the standards for sustainability included in the local Agenda 21 of the Deutsche Städtetag (German Association of Cities), Kronsberg, was set out as an innovation project that would deploy ecological solutions to enhance urban life and sustainable living. The expo-concept clearly focused on energy efficiency optimization, soil management, rainwater management, waste concepts, and environmental communication. Originally a topic of medium importance, rainwater management became one of the central issues as hydrological and technical studies showed that a residential district with standard drainage system in this area would have major impacts on the regional water flows. In order to make construction and development environmentally sound despite this difficult situation, a seminatural drainage concept was developed to minimize the effects of development on the

Water and Sustainable Design

natural water balance and to safeguard infiltration and groundwater refill (Fig. 10). BGI in Portland, Oregon (USA)

Portland is known as one of the most forwardthinking cities in the USA in terms of promoting and advocating sustainability [10]. To start, Portland purchased and permanently protected more than 33 km2 of ecologically valuable natural areas from future development and has continued to show a strong commitment to environmentally conscious land use, including an approach to land conservation and enhancing green areas (Parks Vision 2020). Portland has also emerged as a pioneer in promoting compact city design through municipal policy. In 1996, a Stormwater Policy Advisory Committee (SPAC), stakeholders from landscape architecture, architecture, engineering, institutional organizations, and the stormwater treatment industry were assembled to generate recommendations and guidelines for urban stormwater engineering and design. In 2010, Tanner Springs Park was created through intense community participation to address the challenges of a compromised

Water and Sustainable Design, Fig. 10 Hannover-Kronsberg, Germany, uses ecological BGI solutions to enhance urban and sustainable living. (Photo Credit: Herbert Dreiseitl)

Water and Sustainable Design

neighborhood built on a former wetland and creek bed. Tanner Springs Park transformed the Pearl District through BGI, using stormwater runoff to feed a natural spring and a natural cleansing system. Today, ospreys dive into the water, art performances unfold on the floating deck, children splash and explore, and others take quiet contemplation in this natural refuge in the heart of the city means that this park is the realization of the dreams and hopes of local people (Fig. 11). Since that time, Portland has been recognized as a leader in “green” stormwater management with a number of award-winning BGI projects including the “Portland Ecoroof Program” the “Green Streets” project and a number of pervious pavement projects. Portland’s multi-stakeholder governance structure presents an interesting institutional context in which BGI projects have been successful.

199

centric approach, predicated on access to daylight, ventilation, views, and the presence of gardens and nature. Patient and visitor areas are placed around a landscaped central garden. This garden opens up to an adjacent Yishun stormwater pond from which it taps vistas and breezes. Visitors from nearby housing estates now use the hospital’s public spaces for enjoyment alongside patients and other official visitors. In 2005, KTPH team expanded its blue-green footprint by adopting the adjacent Yishun Pond a gray infrastructure for water storage. The Hospital now has a lush BGI central garden flowing into a waterfront promenade with an energizing walking trail that wraps the lake (Fig. 12). The former gray infrastructure retention pond now gives a picturesque view for pedestrians, as its concrete edge has been softened with planting, and artificial floating wetlands have been added attracting new species.

BGI at Khoo Teck Puat Hospital and Yishun Pond (Singapore)

BGI in Hamburg (Germany)

Khoo Teck Puat Hospital (KTPH) is one of the most recent of seven public hospitals in Singapore. It is set out to widen the perspective on healthcare in Singapore to include healing spaces in which the design of the physical environment actively contributes to wellness. This translated into the integration of biophilic elements. The KTPH design brief spoke explicitly of a patient-

Hamburg is situated on the river Elbe and hosts one of the largest harbors in Europe. Situated only 6 m above sea level and increasingly hit by heavy rainfall, severe flooding and associated damages increasingly threaten central Hamburg. The high density of buildings and imperviousness surfaces increase the risk of flooding and severely challenge the existing rainwater system. In 2009,

Water and Sustainable Design, Fig. 11 Tanner Springs Park transformed the Pearl District in Portland, Oregon, through BGI. (Photo Credit: Herbert Dreiseitl)

200

Water and Sustainable Design

Water and Sustainable Design, Fig. 12 The KTPH hospital in Singapore has a lush BGI central garden flowing into a waterfront walking trail. Khoo Teck Puat

Hospital (KTPH) and Yishun Pond, Singapore. (Photo Credit: Giovanni Cossu (LCL Ramboll))

Hamburg introduced an initiative to develop a rainwater adaptation plan – RISA – in which all relevant agencies (water, park and urban green, traffic, environment) were required to cooperate and develop comprehensive and holistic guidelines for a satisfactory infrastructure intervention. BGI is expected to have a prominent position in the new design, especially since individual, smaller-scale BGI projects (e.g., Kleine Horst in Hamburg Ohlendorf) have proven to be very successful.

attention in Copenhagen since it is a coastal town that is at increased risk from flooding due to the rising sea level combined with increased frequency of extreme precipitation events. Moving to address the increased flooding risks, the Copenhagen Climate Adaptation Plan of October 2011 promoted the incorporation of BGI, especially retention areas, within the urban landscape. Led by Ramboll Studio Dreiseitl, the strategy addresses key issues of flood management and water quality while seeking to create the greatest possible synergy with the urban environment (Fig. 13). A “CloudBurst” tool box of urban interventions, such as CloudBurst boulevards, CloudBurst parks, and CloudBurst plazas, provides the basis for a dynamic and multifunctional system. This new generation of blue-green infrastructures addresses essential city services such as mobility, recreation, health, and biodiversity, creating a strategic and feasible approach to ensure long-term resilience and economic buoyancy.

BGI in Copenhagen (Denmark)

Copenhagen, the capital and most populous city in Denmark, is known internationally as an outstanding example for high livability and futureoriented urban design. Surveys have shown a high degree of public awareness and political support for sustainability- and livability-related issues. Climate adaptation in response to global warming is one of the major topics worthy of special

Water and Sustainable Design

201

Water and Sustainable Design, Fig. 13 A new generation of blue-green infrastructures for Copenhagen addresses essential city services such as mobility,

recreation, health, and biodiversity. VANDPLUS Project info Viborg – sùnës Dispositions plan. (Photo Credit: Ramboll)

Copenhagen is rich in social resources (knowledge, institutional capability, financial capital) that are required in the step-by-step restructuring of the densely populated and builtup inner-city areas – those that have experienced the most frequent and intense flooding. Copenhagen provides an interesting case for examining aspects of political and institutional framing and negotiations for BGI – implementation.

was analyzed as a change in an urban society’s pool of resources for a decent life, according to criteria of livability, sustainability, and resilience. Therefore, all relevant resources are defined as different forms of societal capital: the natural, built, human, social, symbolic, and financial capital. As a consequence, financial capital can be treated equally in the context of all other capitals relevant for quality of life and long-term social development. In our study, the term “capital” is used for all relevant societal resources. While the term capital is usually understood as financial capital, i.e., a final monetizable outcome of economic transactions, the modern understanding of the term has broadened this meaning, applying it more generally to other types of resources used in society. In a nutshell we follow a triple bottom line methodology to take economical, ecological (defined as natural capital), and social sustainability as three pillars that represent distinct dimensions for

Modeling of BGI-Induced Change on Urban Society In order to assess the societal (including ecological and economic) impacts of BGI implementation, the authors modeled the BGI-induced change of an urban society’s capability for livability, sustainability, and resilience. In particular we employed a socioeconomic capital-based accounting model, based on the “Polychrome Sustainability” approach of Manfred Moldaschl. The implementation of BGI in dense urban areas

202

evaluation. However, we suggest applying a more detailed and elaborated version of the social pillar to include human capital, social capital, and symbolic capital. Human, social, and symbolic capitals are types of immaterial capital, a type of capital that is considered to differ crucially from financial capital and natural capital both in their forms of manifestation and in their forms of (re-) production. Immaterial capital may or may not be monetized. The different categories of immaterial capital are inseparably linked to human competences and/or social relations. Immaterial capitals often follow a more generic logic such as trustful behavior is built on trust and enhances trust.

BGI Benefits for Health and Well-Being: Bishan-Ang Mo Kio Park The effects of BGI implementation on human health, public well-being, financial assets, longterm economic resources, and other human values have been identified through the case study and comparative analysis of the Bishan-Ang Mo Kio

Water and Sustainable Design

(BAMK) Park in Singapore. Solving serious flooding and water quality challenges with this park led to a dramatic increase in accessible open space for neighboring communities, with benefits for social life and improved awareness of these communities in ecology and the environment (Fig. 14). The highlight of this project is the revitalization of the river. The unique plan to break the concrete channel and create a naturalized waterway was initiated for the first time in Singapore. Designed on a floodplain concept, people can enjoy recreational activities along the riverbanks during dry weather and during heavy rain; the park land adjacent to the river doubles up as a conveyance channel, increasing carrying capacity by 40%. This enables multiple land uses within the park, creating more spaces for the community as well as ecologically valuable and diverse habitats (Figs. 15 and 16). To date, the park has seen the park’s biodiversity increase by 30% with 66 species of wildflower, 59 species of birds, and 22 species of dragonfly identified – with some being identified as rare in a city environment.

Water and Sustainable Design, Fig. 14 (BAMK) Park in Singapore led to a dramatic increase in accessible open space for neighboring communities; Tai Chi in Bishan-Ang Mo Kio Park, Singapore. (Photo Credit: Herbert Dreiseitl)

Water and Sustainable Design

203

Water and Sustainable Design, Fig. 15 The old concrete canal in Bishan-Ang Mo Kio Park

Water and Sustainable Design, Fig. 16 A unique BGI plan to replace the concrete channel and create a naturalized waterway in Bishan-Ang Mo Kio Park, Singapore. (Photo Credits: Herbert Dreiseitl)

As a result of the redevelopment of BAMK into a naturalized park, the number of park visits has doubled from 3 to 6 million persons/year (Fig. 17). It was found that after the BGI upgrade to BAMK, nearly 50% of all park users were engaging in active physical activities, such as jogging, bicycling, skating, or intense walking. It has been estimated that the positive impact on physical health for the community is substantial, estimated at SGD 16–43 million (which is 12–31 million USD at 2013

exchange rates). Moreover, the researchers also identified mental health benefits attributable to BAMK’s ability to attract social life and to encourage social integration. The combination of natural beauty and the dual physical assets of a park and a river appeals to people. As they get close to the water and appreciate nature’s rhythms and wonder, their experience of water and their sense of responsibility to their environment change, leading to collective goals to be better stewards of the environment.

204

Water and Sustainable Design

Water and Sustainable Design, Fig. 17 The number of park visits has doubled from 3 to 6 million persons/year in Bishan-Ang Mo Kio, Park, Singapore. (Photo Credit: Herbert Dreiseitl)

Conclusion: Water’s Position in Sustainable Urbanism

Ten Guidelines for Working with Water in Sustainable Urban Design

History will record that the political and design leaders of the past narrowly reduced the role of water to the most basic technical and engineering services, as liabilities and not assets. Yet, water has always been a critical driver in shaping cities, defining location, and creating character and atmosphere [11]. After generations of industries have used rivers as dumping grounds, cities worldwide are now turning their faces to the water and looking for ways to rejuvenate our natural resources. Greenblue infrastructures are becoming more and more important as a dynamic resource to balance and stabilize life processes and as a backbone for livability. Water has a strong positive biophilic effect on people and defines the symbolic capitals of cities like Amsterdam, Sydney, New York, Venice, and Pittsburgh just to mention a few. There are a number of critical guidelines for bringing all of the regions’ water resources into a vision for the future:

1. Think fluid, resilient, and regenerative; water is the best teacher to have the right attitude in design, and technical and aesthetic solutions have to be in line with the nature of this most dynamic and lively substance. 2. Follow and search for multifunctional solutions; water is always interacting with its surroundings and has a strong impact on air, temperature, energy, metabolic systems, as well as flora and fauna. 3. Use water as a balancing regulator to reduce temperature extremes and reduce dust and sound pollution; combine blue-green infrastructure to filter; and avoid environmental stressors and shocks such as climate change and sea level rise. 4. Combine energy with water-related systems; too much energy is used for pumping and processing water; and decentralization and gravity-based solutions should be prioritized to reduce significant energy consumption.

Water and Sustainable Design

5. Work with regenerative systems and food production; more than 70% of the world’s water is used for food production and processing, and we have to implement systems that are regenerative and local nature-based solutions. 6. Reduce water consumption to a minimum and reuse water in cascading system; balance systems between local and regional decentralized systems and larger, centralized systems. 7. Avoid and upcycle waste in the metabolic process of water; water is most stressed by waste, so avoid using water as a transport medium for waste, finding better solutions for recycling and upcycling waste. 8. Ensure access to water for all; equity and social justice need to be respected in waterrelated planning process; and involve stakeholders and the public as a means to care for social equity. 9. Bring beauty in all water design, as water is intimately connected to beauty; respect the human fascination with water and the biophilia effect, which we so much need today. Water is key to our survival in cities and rural areas. It is critical to measure, count, and quantify the impacts of water design, capturing the value and character of water’s flexible and everchanging qualities. The future depends on the design of our water and landscape systems as a central tenet of a sustainable built environment.

205

Bibliography 1. Dreiseitl H, Grau D (2009) Recent waterscapes planning, building and designing with water. Birkhäuser, Basel, pp 74–78. ISBN 978-3-7643-8984-0 2. Schröpfer T (2015) Dense+ green: innovative building types for sustainable urban architecture. Birkhäuser, Basel, pp 48–60. ISBN 978-3-03812-579-0 3. Cities as Living Systems (2016) Urban solutions, 8. Centre of Liveable Cities, Singapore, pp 14–21, Interview with Dreiseitl. H. Retrieved from https:// www.clc.gov.sg/docs/default-source/urban-solutions/ urban-solutions-8-full.pdf 4. Baykal A (2012) Cloudburst Management Plan 2012, The city of Copenhagen Technical and Environmental Administration, Copenhagen pp 1–28 5. President’s Design Award Singapore, Design of the Year 2012, Bishan-Ang Mo Kio Park. (2012). Retrieved Feb 2019, from https://www.designsingapore.org/presidentsdesign-award/award-recipients/2012/bishan-ang-mo-kiopark.html 6. Dreiseitl H, Wanschura B (2014) Strengthening bluegreen infrastructures in our cities. Ramboll, Liveable Cities Lab 7. Liptan TW (2017) Sustainable stormwater management: a landscape-driven approach to planning and design. Timber Press, Portland, Oregon 8. Wörlen M, Wanschura B, Dreiseitl H, Noiva K, Wescoat J, Moldaschl M (2016) Enhancing bluegreen infrastructure and social performance in high density urban environments: summary document. Ramboll Liveable Cities Lab, Uberlingen 9. Yudelson J (2012) Green building trends: Europe. Island Press, Washington, DC, p 23. ISBN 978-1597264778 10. LOFT Publications (2009) Sketch landscape. Page One Publishing, Singapore, pp 270–291. ISN 978-981-245-738-7 11. Dreiseitl H (2015) Liveable cities; the art of integrating today what we need tomorrow. FuturArc J 38:22–31. BCI Asia

Green Infrastructures to Face Climate Change in an Urbanizing World Stephan Pauleit1, Ole Fryd2, Antje Backhaus3 and Marina Bergen Jensen2 1 Center of Life and Food Sciences Weihenstephan, Technical University of Munich, Freising, Germany 2 Landscape Architecture and Planning, Department of Geosciences and Natural Resource Management, University of Copenhagen, Frederiksberg C, Denmark 3 gruppe F, Berlin, Germany

Article Outline Glossary Definition of the Subject Introduction The Role of Green Infrastructure for Adaptation of Cities to Climate Change The Potential of Green Infrastructure to Mitigate Climate Change Vulnerability of the Green Infrastructure to Climate Change Impacts Conclusions and Future Directions Bibliography

Glossary Adaptation (with respect to climate change) “The process of adjustment to actual or expected climate and its effects. In human systems, adaptation seeks to moderate or avoid harm or exploit beneficial opportunities. In some natural systems, human intervention may facilitate adjustment to expected climate and its effects” ([1], p. 5). A distinction has been made between planned adaptation (e.g., urban planning), which is the focus of this chapter, and autonomous adaptation (e.g., by

individual action such as improving housing insulation, installing air-conditioning, etc.) [2]. Ecosystem services “The benefits people obtain from ecosystems” [3]. “These include provisioning services such as food, water, timber, and fiber; regulating services that affect climate, floods, disease, waste, and water quality; cultural services that provide recreational, aesthetic, and spiritual benefits; and supporting services such as soil formation, photosynthesis, and nutrient cycling” ([3], Preface: V). In urban areas, ecosystem services are clearly related to land use and land cover. Therefore, spatial planning and regulations that influence the spatial pattern and intensity of land use, and in particular the provision and quality of green spaces, can have huge implications for the ecology of cities (e.g., [4, 5]). Evapotranspiration The sum of evaporation of water from surfaces and the transpiration of water by plants and animals. According to the US Geological Survey [6], transpiration from plants accounts for approximately 10% of air moisture. A large oak has been estimated to transpire up to 151,000 liters of water per year [7]. This would account for more than 400 l/day. More modest figures are given by other sources for trees in urban streets and squares, e.g., around 50 l/day for 30–40-year-old narrow-leaved lime trees, but only around 13 l/day for black locust trees, in Munich, Germany [8]. In situ measurements of 108 urban trees in Los Angeles showed evapotranspiration rates ranging from close to zero to 200 l/day with high variability among tree species and locations [9], as well as irrigation patterns, air moisture, and exposure to sunlight [10]. Furthermore, the health and growth may vary for each individual urban tree. Green Infrastructure The term “green infrastructure” was first introduced in the USA in the 1990s – defined as an “interconnected network of protected land and water that supports native species, maintains natural ecological processes, sustains air and water resources, and con-

© Springer Science+Business Media, LLC, part of Springer Nature 2020 V. Loftness (ed.), Sustainable Built Environments, https://doi.org/10.1007/978-1-0716-0684-1_212 Originally published in R. A. Meyers (ed.), Encyclopedia of Sustainability Science and Technology, © Springer Science+Business Media, LLC, part of Springer Nature 2020, https://doi.org/10.1007/978-1-4939-2493-6_212-3

207

208

Green Infrastructures to Face Climate Change in an Urbanizing World

tributes to the health and quality of life for America’s communities and people” [11]. “Urban green infrastructure” is the network of green areas in cities. The term makes reference to other types of urban infrastructures (e.g., the road system). This interpretation of green infrastructure relates to a fine-scale urban application where hybrid infrastructures of green spaces and built systems are planned and designed to support multiple ecosystem services. It has been argued that planning of an urban green infrastructure should promote multifunctionality and connectivity of urban green space. It can integrate both public and private green space and seek to develop synergies with technical and social infrastructures. It should be based on a long-term vision and a communicative and socially inclusive approach to its planning and management [12, 13]. Mitigation (with respect to climate change) Reducing greenhouse gas emissions and enhancing sinks. Resilience “The capacity of social, economic, and environmental systems to cope with a hazardous event or trend or disturbance, responding or reorganizing in ways that maintain their essential function, identity, and structure, while also maintaining the capacity for adaptation, learning, and transformation” ([1], p. 5., see also [14, 15]). Resilient systems can be characterized by qualities such as diversity, modularity, feedback, and redundancy [16]. Despite numerous definitions of resilience [17], it is relevant to highlight Davoudi’s [18] distinction between engineering resilience focusing on the pace of a system to “bounce back” to the preceding condition after a disturbance, often with an emphasis on technological measures to protect against a disturbances of the status quo; ecological resilience emphasizing the ability of the intertwined social, ecological, and technological system [19] to adapt and adjust to changing internal and external processes, while maintaining the existing regime; and evolutionary resilience highlighting the ability of a system to “bounce forward” along sustainability transition pathways away from a less desired current regime to a more desired future regime which requires deep systemic change [20].

Sustainable Urban Drainage Systems Sequence of management practices utilizing urban green areas for storage, infiltration, evaporation, and conveyance of stormwater runoff. Urban Heat Island Effect Significantly higher temperatures experienced in cities compared to the rural surroundings as a result of poor solar reflection in the built environment, less evapotranspiration, and anthropogenic heat from combustion engines, conditioning of buildings, and other use of energy. Vulnerability (with respect to climate change) It is “the propensity or predisposition to be adversely affected. Vulnerability encompasses a variety of concepts and elements including sensitivity or susceptibility to harm and lack of capacity to cope and adapt” ( [1], p. 5). Vulnerability is determined by a broad set of factors including physical properties of the system in focus, e.g., construction of buildings, preparedness measures, environmental management [21], wealth, social status, and gender, as well as the adaptive capacity of individuals, communities, and institutions [22].

Definition of the Subject By 2050, more than two thirds of the world’s population will live in cities [23]. Already today, cities of the developed world are a major source of greenhouse gas emissions. Therefore, cities need to make serious efforts to mitigate climate change. Urban planning can play a major role in this respect by designing compact cities with a low carbon footprint. At the same time, climate change is having a severe impact on cities with the intensification of the heat island effect and the increase of surface runoff from more frequent and intense rainstorms and by coastal and river flooding. Urban planning will play an important role for development and implementation of integrated strategies for climate change mitigation and adaptation. Most critically, “green infrastructures” can assist in adapting cities to climate change by reducing the urban heat island effect and by managing stormwater runoff. Urban greening such as planting of shade trees and roof greening can also reduce the energy demand for house heating and cooling. The design of the urban

Green Infrastructures to Face Climate Change in an Urbanizing World

landscape will have a direct influence on mitigating climate change and its impact on people’s livelihoods and assets.

Introduction Urbanization and climate change are closely related. At present, more than half of the world’s population lives in cities. Over the next 30 years, the world’s urban population is projected to increase by approximately 2.5 billion people [23]. This will increase the urban population by 60% and lead to massive land use changes in and around current urban settlements. Close to 90% of the future urban expansion is expected to take place in Asia and Africa [23]. While new urbanization is an issue in developing countries, 70–90% of the population of the developed world already lives in urban settlements. Therefore, the importance of green infrastructures for climate change mitigation and adaptation is an equally important issue in the developed as well as the developing world. It has been estimated that 78% of carbon emissions from burning fossil fuels and cement manufacturing [24] and 85% of the anthropogenic emissions of carbon dioxide, chlorine-fluorinecarbons, and tropospheric ozone stem from urban areas [25]. In another study, the contribution of urban areas to global greenhouse gas emissions has been estimated between 37% and 49% for the year 2000 [26]. While these and similar figures have been debated [27], there can be little doubt that urban areas, in particular in the developed world and in the transition countries, need to play a major role in climate change mitigation. Accordingly, the abatement of greenhouse gas emissions is rising in the political agenda of many cities, in particular in the developed world [28–30]. Greenhouse gas emissions can be reduced in many different ways by all sectors and at all levels of human society. Urban form plays a critical role, and the “compact city” model has gained wide acceptance as a way to achieve more resource efficient cities. Walkability, mixed-use, and urban density are strongly related to energy consumption, with denser cities consuming less energy per capita. This applies especially for car-based travel [31]. This relationship has been used as a powerful argument

209

against allowing further urban sprawl, i.e., the extension of urban areas through low-density developments. European cities have on average expanded by 78% in area since the mid-1950s, while their population increased by only 33% [32]. Large urban regions have developed, far into the previous countryside, and have resulted in patterns of daily commuting [33]. Sprawl has even been observed in city regions with decreasing populations where people move out from declining inner cities to live in suburban and even rural locations [34]. While city governments are giving increasing attention to reducing greenhouse gas emissions for climate mitigation, adaptation to climate change has been of less concern until recently. A study of 627 urban climate change initiatives in 100 cities found that only 12% of the initiatives were addressing climate change adaptation [35]. Another study of 401 large cities with a population exceeding one million people found that only 15% of the cities had climate change adaptation initiatives [36]. A similar trend is seen in a review of local climate plans in 885 cities across the European Union where approximately 25% of the cities had adaptation plans while 42% had mitigation plans and 33% of the cities had no stand-alone local climate plan [30]. Following the IPCC Fifth Assessment Report [1, 37] and Crichton [38] (Fig. 1), the climate change-related risks of urban areas are a function of three factors: the (i) hazards, (ii) exposure of the urban system to these hazards, and (iii) the urban system’s inherent vulnerability. (i) Hazards: The average global mean surface temperature is likely to rise between 0.3  C and 4.8  C, and the sea level is likely to rise between 0.26 m and 0.82 m by 2081–2100 relative to 1986–2005 – depending on the “Representative Concentration Pathway scenario” (IPCC Fifth Assessment Report [39]). Annual precipitation levels will increase in high latitudes and decrease in most subtropical land regions. Temperature extremes, heat waves, and heavy rain events are expected to increase [1]. Overall, three effects of climate change are of particular concern for urban areas: • Sea-level rise and increase of storm surges caused by storms

210

Green Infrastructures to Face Climate Change in an Urbanizing World

Green Infrastructures to Face Climate Change in an Urbanizing World, Fig. 1 Illustration of risk as a function of hazard, exposure, and vulnerability ([37], see also

[1]). (Adapted from Crichton (2001) by Handley [169] with permission from author)

• Temperature rise which intensifies the urban heat island • Changing amounts and patterns of precipitation which increases the risk of drought, on the one hand, and pluvial and sewer flooding and landslides from rainstorms, on the other hand Temperatures will increase in particular in the North and the South of Europe, but in more moderate ways in Central and North West Europe (Fig. 2). In the North, temperatures and precipitation will particularly rise in winter time, while the South will be affected by hotter and drier summers. Not all of these changes will be negative for living in the city. For instance, heating demands and human winter mortality are expected to decrease in northerly areas. Also, the summer will become more beneficial for outdoor activities in the north. Yet, even for countries such as Sweden, it has been estimated that these beneficial effects are outweighed by increased mortality due to heat waves and increased energy demands for cooling during summertime [40]. In particular, it needs to be highlighted that it is less

the general change of temperature and precipitation patterns that are of concern, but the increasing frequency and intensity of extreme events such as spells of hot weather and drought interspersed with more frequent thunderstorms which bring large quantities of rainfall in a short time and a higher likelihood of more frequent and heavier winter storms in the northern, northwestern, and central parts of Europe [41]. (ii) Exposure: Urban areas are often located in zones particularly exposed to climate change hazards such as storm surges, river floods, and landslides. About 13% of the world’s urban population live in the low-elevation coastal zone (