Life Cycle Design: An Experimental Tool for Designers [1st ed.] 978-3-030-11496-1, 978-3-030-11497-8

This book proposes an economic and environmental assessment tool to help private and public building designers and owner

378 104 5MB

English Pages XIII, 157 [170] Year 2019

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Life Cycle Design: An Experimental Tool for Designers [1st ed.]
 978-3-030-11496-1, 978-3-030-11497-8

Table of contents :
Front Matter ....Pages i-xiii
Life Cycle and Sustainability: Concepts and Keywords (Francesca Thiébat)....Pages 1-20
The Environment and Economics (Francesca Thiébat)....Pages 21-30
Life Cycle Methodologies (Francesca Thiébat)....Pages 31-64
Defining an Innovative Design Method Based on the Life Cycle Approach (Francesca Thiébat)....Pages 65-118
Case Studies (Francesca Thiébat)....Pages 119-151
Conclusions and Outlook (Francesca Thiébat)....Pages 153-157

Citation preview

PoliTO Springer Series

Francesca Thiébat

Life Cycle Design An Experimental Tool for Designers

PoliTO Springer Series Series editors Giovanni Ghione, Turin, Italy Pietro Asinari, Department of Energy, Politecnico di Torino, Turin, Italy Luca Ridolfi, Turin, Italy Erasmo Carrera, Department of Mechanical and Aerospace Engineering, Politecnico di Torino, Turin, Italy Claudio Canuto, Department of Mathematical Sciences, Politecnico di Torino, Turin, Italy Felice Iazzi, Department of Applied Science and Technology, Politecnico di Torino, Turin, Italy Andrea Acquaviva, Informatica e Automatica, Politecnico di Torino, Turin, Italy

Springer, in cooperation with Politecnico di Torino, publishes the PoliTO Springer Series. This co-branded series of publications includes works by authors and volume editors mainly affiliated with Politecnico di Torino and covers academic and professional topics in the following areas: Mathematics and Statistics, Chemistry and Physical Sciences, Computer Science, All fields of Engineering. Interdisciplinary contributions combining the above areas are also welcome. The series will consist of lecture notes, research monographs, and briefs. Lectures notes are meant to provide quick information on research advances and may be based e.g. on summer schools or intensive courses on topics of current research, while SpringerBriefs are intended as concise summaries of cutting-edge research and its practical applications. The PoliTO Springer Series will promote international authorship, and addresses a global readership of scholars, students, researchers, professionals and policymakers.

More information about this series at http://www.springer.com/series/13890

Francesca Thiébat

Life Cycle Design An Experimental Tool for Designers

123

Francesca Thiébat Dipartimento di Architettura e Design (DAD) Politecnico di Torino Turin, Italy

ISSN 2509-6796 ISSN 2509-7024 (electronic) PoliTO Springer Series ISBN 978-3-030-11496-1 ISBN 978-3-030-11497-8 (eBook) https://doi.org/10.1007/978-3-030-11497-8 Library of Congress Control Number: 2018968090 © Springer Nature Switzerland AG 2019 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, express or implied, with respect to the material contained herein or for any errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. This Springer imprint is published by the registered company Springer Nature Switzerland AG The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland

Foreword

They say, maybe indulging in the paradox, that Marcus Vitruvius Pollio was the first to have the intuition of the “sustainable development” concept; not so much for the famous three word punch-line “firmitas, utilitas, venustas” as for the more suggestive image of the Genius Loci, the Spirit of the Place. The ineffable, ubiquitous, austere Lord of the Place to whom we owe respect when, with the violence of our artefacts, we trespass into his territory. The Lord of the Place was not a creature by Vitruvius, but an idea with deep roots in the Roman lay and religious culture. It was an ideal concept coming from faraway places, the Asian deserts, Banda Sea Islands, Sunda Sea, wherever—places where the ineffable Lord is still alive and treated with due reverence. Every day, at the appointed times, set by old traditions, beautiful girls, young ladies and old women gracefully bring offerings of flowers, fruits and leaves to the Spirits of the places to show their respect and to convey a silent prayer for benevolent protection. It is a small rite with deep meaning: we are guests in these places which belong to others and, as temporary guests, we must behave. It may not be by chance that the modern idea of “sustainability” came to a Norwegian woman (Gro Harlem Brundtland 1987) living in a country with long freezing Winters and cool short Summers. Let us use the environment in such a way that will allow future generations to use it as well, she said, more than 30 years ago. We are guests and as guests we should behave. Gro Harlem did not invent the concept, which is the modern interpretation of the Vitruvian Genius Loci that the Romans learned from the Etruscans, who probably got it from the Phoenicians, who heard it from merchants/travellers coming from Asia— passing like a prehistoric Olympic torch through centuries, peoples, oceans, deserts, mountains, arctic cold and tropical heat. Now it is here with us after a few hundred years of environmental massacres, monumental squandering, wars, waste, poisons, middle ages, renaissances, industrializations, globalizations, ideological utopias and Cartesian rationalities—perhaps beyond the point of no return. There is no possible denial, no return to a sender who does not exist. We have to welcome it, accept it, foster, manage, respect and implement it. We, the sons of wild squandering, of a culture built on the arrogant illusion of infinite resources and infinite environmental space to waste, of an economy where the cost of the environment is zero, nothing, v

vi

Foreword

zilch… and the future generations: stuff them all! We, the careless, who live in and design lethal, suicidal cities where nobody knows how others will be able to live, cities that somebody before us designed and built, somebody who could not care less about how we could live in them… Yes, it is us, just us the ones who have to welcome the young, delicate child of sustainability, listen to him, understand, translate, implement him, feed him and raise him to be the Powerful Warrior, Prince, Master and Commander who will manage the planetary intergenerational processes to change the whole Planet in the coming century. An Architect of a kind we have never known or seen before. A huge responsibility and commitment. No prisoners taken, a slim chance of success. No certainty granted. Because the sustainable future of the Planet is by far the toughest revolution ever attempted by humankind. At stake nothing less but the future of the Species. It is impossible to deny; it is necessary and we must face the challenge. And here comes Francesca Thiébat’s book, thirty-two years after Gro Harlem Brundtland’s intuition, twenty-something centuries after Vitruvius and a few millennia after the Asian pervasive Gods of the Place. The book guides through the complexity of design considering the whole life cycle assessment of materials, components, buildings and towns in which not only we live, but also future generations will live. Remarkable feature of this book is the attention to environmental costing: a matter not usually dealt with in current sustainability literature. Francesca Thiébat designs, prepares new tools, sets them out, grammar, methods, know-how and then the synopsis of the Complex Sustainable Architectural Design. The very core of our Discipline: to know-how to do things and make them happen. The book is a huge step forward on the long and difficult road from Gro Harlem’s intuition, from the Vitruvian Genius, from the lovely pagan Asian Deities, the road that, beyond Renaissance, will lead us to the knowledge of doing things and make them happen. A huge step forward for the child who has trusted us, now a powerful Kind Warrior: the warrior of Francesca Thiébat. Turin, Italy

Prof. Gabriella Peretti

Acknowledgements

First of all, I would like to thank my parents. My mother, for having pushed me towards the more experimental, creative and solar side of life, teaching me to measure myself against freedom, progress and ethics. My father, for introducing me to the potential of three-dimensional space and teaching me it can be turned into architecture. My thanks go to the colleagues at the Department of Architecture and Design of the Politecnico di Torino with whom I have worked and experimented. Gabriella Peretti for instilled in me the technique and passion required to organise and perform research, for having believed in my work from the very start, for encouraging me to seize the opportunity to write the book and for her invaluable suggestions regarding the text. Lorenzo Matteoli for sharing his thoughts on sustainability, technology, macroeconomics and architecture. Daniela Bosia for always supporting and involving me in several projects on my research topic and Life Cycle Assessment (LCA). All these work experiences have allowed me to fine-tune my skills in this area. I would also like to thank Roberto Giordano for our long conversations over the years, for exchanging information about environmental design and for encouraging me to tackle the innovative and somewhat daring theme of Life Cycle Costing associated with LCA. Corrado Carbonaro for sharing many research projects on LCA issues; Lorenzo Savio, Roberto Pennacchio, Andrea Levra Levron and Valentino Manni for the experiments on materials and components; Riccardo Pollo and Donatella Marino for the information on maintenance and end of life. Special thanks go to: Rete Italiana LCA for having repeatedly involved me in multidisciplinary teaching experiences on the issues of environmental Life Cycle Costing; Maria Chiara Torricelli for her educational input on the Life Cycle Thinking integrated with the Technology of Architecture; and the Turin Fondazione per l’Ambiente. I would like to also thank the students who with their design applications participated in the improvement of the model.

vii

viii

Acknowledgements

The PAT architecture firm for supporting my decision to integrate research and practice and for sharing data and images regarding the case studies presented in this book. Alessandro, Michele, Stefano and Giuseppe for their advice, suggestions and critiques of the application of the LCDM model to real cases. Andrea Veglia for helping me find the fil rouge between practice and research which I first used to write my PhD thesis and then this book. My thanks also go to my friends and relatives who have inspired my ideas and allowed them to flourish. I equally extend my thanks to those who made me trip and fall in life. Those obstacles have made me stronger and more courageous. Finally, I must thank my daughters; they allowed me to dedicate time to my work which I believe is part of my life. I dedicate this book to them, a book that speaks of the future, of the past (without which there would be no future) and of the fil rouge that running between them. I hope that their future will also be experimental, creative and solar. Part of the research on which this book is based was conducted, thanks to the contribution of the European Social Fund (ESF) and the Valle d'Aosta Region (2006–2008). The collection and processing of inventory data of the Chavonne Warehouse case study is to be attributed to Valentina Porceddu; the graphic elaboration of the diagrams contained in chapters 1, 2, 3 and 4 (excluding: 4.9 and figures from 4.14 to 4.18) to Chiara de Grandi, Fig. 4.9 to Chiara Rota, Fig. 5.2 to Alberto Matta, Fig. 5.6 to BuonomoVeglia srl and Fig. 5.12 to Nicolò Radicioni. All the photographs and drawings of the case studies were provided by PAT. All other diagrams have been elaborated by the author.

Contents

..... .....

1 1

..... ..... .....

5 9 10

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

10 12 14 16 19

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

21 21 23 24 27 28 29

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

31 31 31 35 42 42 49

1 Life Cycle and Sustainability: Concepts and Keywords . . . . . 1.1 Architecture/Ecology/Sustainable Development . . . . . . . . . . 1.2 Design Process/Multidisciplinary Approach/Life Cycle Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3 Time/Service Life/Future . . . . . . . . . . . . . . . . . . . . . . . . . . 1.4 Regulations/Standards/Codes . . . . . . . . . . . . . . . . . . . . . . . 1.4.1 ISO/TC 59 Buildings and Civil Engineering Works (SC14 and SC17) . . . . . . . . . . . . . . . . . . . . . . . . . 1.4.2 CEN/TC 350 Sustainability of Construction Works . 1.5 Integrated Design/Assessment Systems/Tools . . . . . . . . . . . 1.5.1 Tools for Integrated Design . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 The Environment and Economics . . . . . . . . . . . . . . . 2.1 The Circular Economy . . . . . . . . . . . . . . . . . . . . 2.1.1 The Carrying Capacity of the Environment 2.2 Environmental Economics . . . . . . . . . . . . . . . . . . 2.3 Tools and Principles of the Environmental Policy . 2.4 The Circular Approach to Design . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

3 Life Cycle Methodologies . . . . . . . . . . . . . . . . 3.1 Life Cycle Assessment (LCA) . . . . . . . . . . 3.1.1 Origins, Definitions, and References 3.1.2 Environmental Indicators . . . . . . . . 3.2 Life Cycle Costing . . . . . . . . . . . . . . . . . . 3.2.1 Origins, Definitions and References 3.2.2 General Framework and Costs . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

ix

x

Contents

3.3 Combining LCA and LCC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.1 A Common Framework for the Building Sector . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 Defining an Innovative Design Method Based on the Life Cycle Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1 Design with a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 The Life Cycle Design Model (LCDM) . . . . . . . . . . . . . . . . . 4.2.1 Basic Principles of the Life Cycle Design Model . . . . . 4.2.2 General Framework . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.3 The Environment and Energy Factor . . . . . . . . . . . . . . 4.2.4 The Economic Factor . . . . . . . . . . . . . . . . . . . . . . . . . 4.3 €CO: Economic-Environmental Efficiency Factor . . . . . . . . . . 4.3.1 Example of How to Calculate the Efficiency Factor . . . 4.4 Interpretation of the Results . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 Case Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1 Life Cycle Design for Building Envelopes . . . . . . . . . 5.1.1 Building Façades: VM House . . . . . . . . . . . . . 5.1.2 Green Roof: TEA Headquarters . . . . . . . . . . . 5.2 Life Cycle Design for the Whole Building . . . . . . . . . 5.2.1 Industrial Building Restoration: The Chavonne Warehouse . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

54 56 62

. . . . . . . . . . .

. . . . . . . . . . .

. 65 . 65 . 70 . 73 . 76 . 81 . 84 . 88 . 90 . 93 . 117

. . . . .

. . . . .

. . . . .

119 119 120 128 139

. . . . . . . . . 139 . . . . . . . . . 151

6 Conclusions and Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

Introduction

How can sustainability be fully integrated into architecture in today’s world? Sustainable architecture is characterised by its relationship with the physical and anthropised environment; its objective is to ensure user wellbeing coupled with a reduced consumption of environmental resources and low level of pollution. Eco-compatible design based on Life Cycle Thinking means making choices that have environmental and economic effects, not only during the construction of buildings, but also as regards their management, maintenance and deconstruction. The vision of the actors involved in the project thus requires a “long-term” approach focusing on several strategic objectives, such as: • • • •

minimising the use of natural resources; maintaining efficiency for a pre-established period of time (durability); ensuring adaptability to changes in use over a period of time (flexibility); ensuring the deconstruction and recycling of building components.

A life cycle design methodology should not be applied only to building techniques, materials and energy systems, but above all include the whole design choice process. In fact, although techniques and systems satisfy specific requirements to achieve a reduced environmental impact, when they are combined in a complex structure such as an architectural artefact, then ad hoc parameters need to be used. This holistic approach considers the architecture, the structure, the plant and building systems as parts of an organic whole capable of reciprocally influencing and developing positive synergies, rather than discrete parts to be assessed separately. This outcome can be obtained only by using robustly integrated design. Bearing this in mind, Life Cycle Design not only satisfies the user’s complex requirements, but also promotes sustainable development vis-à-vis three important fields: the economy, the environment and society. This monograph presents the complete results of the author’s research on Life Cycle Design. The study began with her Research Doctorate and has continued to the present day. The experimental model initially developed for her doctoral

xi

xii

Introduction

dissertation was later completed and tested in design projects and real case studies. The final re-elaboration of the model illustrated in this book reflects the ongoing scientific, professional and educational debate which over the years has taken place during research programmes, courses, projects, international congresses and published in Italian and international scientific journals. The first part of the book (Chapters 1, 2 and 3) defines the relationship between environmental and economic aspects in a sustainable design approach. It also illustrates how the life cycle methodology, including Life Cycle Assessment and Life Cycle Costing, can be applied to life cycle design. It presents methods for calculating costs from LCA data, taking into consideration not only a discounted cash flow, but also external costs. It introduces the multidisciplinary approach concept characterising the life cycle design and illustrates the most important internationally recognised tools and theories in the field of architecture. The first part of the book presents the cultural and scientific background for the experimental part of the monograph illustrated in the following chapters. Chapter Four of the monograph presents the experimental design model, called the Life Cycle Design Model (LCDM), based on a life cycle design approach that can be used to produce two different outcomes based on two assessment levels. The first assessment level involves creating a grid, called a Design Matrix, which is useful in the design process. The matrix includes a number of design parameters such as, amongst others, the costs and environmental/energy impacts of a building or building component for each life cycle stage. The second assessment level consists in exploiting LCA and LCC results to develop a user-friendly tool for designers and other actors involved in the building process so that they can assess the most sustainable design option using €CO, a factor to combine the environmental and energy effects of the building system with time and costs. The LCDM allows designers to benchmark their project values against those mandatory by law; it allows them not only to have a clear picture of a life cycle scenario during both the design and use phase of the building when the actors involved in the process will change, but also to compare several design options and implement and scale data quality towards the life cycle stages (from the conceptual design level to monitoring). Furthermore, the double assessments, expressed by the matrix and €CO factor, are also helpful when talking to non-expert clients and communities. Chapter Five uses several case studies to describe the practical application of the Life Cycle Design Model and reveals how environmental impacts and costs can improve the sustainability of a building by extending the building time horizon to the whole life cycle (including construction, maintenance, renovation and end of life). In today’s world, architectural design must necessarily include the three dimensions of sustainability within the ideation process as the tangible expression of the creative and research process. Architects are required to tackle this complex process in a responsible, knowledgeable manner; this process includes the whole life cycle of the building and the effects the built environment has on the environment and those who live in it. Richard Rogers (1992, p. 60) stated: “it is not popular to link the economy and consumption with culture, and to suggest that

Introduction

xiii

today it is the accounting system that dictates the Arts. Yet I firmly believe that to achieve a new cultural enlightenment, one which includes architecture, it will be necessary to redefine the balance between capital, labour, the planet and its poor”. This monograph is addressed to designers and students; it will provide a tool to help not only architects and engineers to define and communicate the principles of sustainability and quality of projects, but also public and private clients to outline their requirements as well as verify and control the design quality and service life of the building.

Reference Rogers R (1992) Architecture: a modern view. Thames and Hudson, New York

Chapter 1

Life Cycle and Sustainability: Concepts and Keywords

Abstract This chapter focuses on the link between the life cycle design approach and the concept of “sustainability”. By reviewing specific keywords it will define the principle of sustainable life cycle design. Starting with this concept the chapter will investigate the state of the art and the role in the design process of regulations, laws, environmental protocols, integrated design tools and software, and the assessment of sustainability in construction. It takes into account both the international scenario and its application in Italy.

1.1 Architecture/Ecology/Sustainable Development How is it possible to fully express the meaning of “sustainable life cycle design”? One definition embracing most of the concepts presented in this book is Dominique Gauzin-Muller’s interpretation of sustainable architecture. She defines ‘sustainable architecture’ as “a balance between rediscovering bioclimatic principles, building traditions emerging from the context, and ingenious innovations that diminish resource use” (Gauzin-Muller 2012, p. 10). This definition contains several key concepts including: the importance of climate in design, the precious legacy of earlier building techniques, close links with the site, technological innovation, reduced resource use, and protection of the ecosystem. These concepts have inspired the adoption of two strategies to achieve sustainability: on the one hand respect for the ecosystem as an ensemble of living and non-living organisms and, on the other, a focus on the time factor, especially the link between past, present, and future. One of the main objectives of ecological, eco-efficient, or sustainable architecture is to curb the environmental impact of the built by reducing resource consumption and pollution throughout the life cycle. All and any form of transformation of the built environment should, by definition, be ecological, precisely because it structures the relationship between man and his life environment over a period of time (Giacchetta and Magliocco 2007; Steel 2005). The term “ecology” or Ökologie in German (from the Greek oikos, house, dwelling and logos, discourse) was introduced by the evolutionist biologist Ernst H. Haeckel in 1866. It indicates the “science of the relationships between living organisms and © Springer Nature Switzerland AG 2019 F. Thiébat, Life Cycle Design, PoliTO Springer Series, https://doi.org/10.1007/978-3-030-11497-8_1

1

2

1 Life Cycle and Sustainability: Concepts and Keywords

the environment where they live and grow”. At first the term was not widely known or used. In fact, one hundred years later, it still appeared unfamiliar. A conversation in 1962 between the poet Saint-John Perse and his translator Friedhelm Kemp is proof of its status as a relatively unknown term. When Kemp asked about the word écologie: “term unfound. What does it mean?”, Saint-John Perse answered: “his [man’s] natural environment, its original context (with all its components). The word is defined in the Petit Larousse as: ‘the biological relationship of living beings with their natural environment’” (Battaglia 2006, p. 39). One should not forget that the prefix “eco-” was initially associated exclusively with biology and the observation of nature and its original balances (e.g., eco-genetic, ecoide, ecology, ecofilia, etc.). Only later was it used in fields involving man’s actions on and responsibilities towards the environment (e.g., eco-compatible, ecocide, etc.). The term ecofriendly (or eco-compatible in Italian) began to appear in dictionaries in the last decade of the twentieth century.1 In architecture, the adjective ecocompatibile refers to processes or products capable of merging with the environment where man lives and in general with the immediate ecosystem. This indicates that the concept of the environment, which up until the seventies was considered only as the natural environment, now includes the “system of structural interrelationships between a person and his/her immediate surroundings” (Terzi 2001). Before industrialisation, the mechanisms behind the relationship between our human society and the natural environment were based on the functioning of the ecosystem as a self-organised structure that produces order in the form of organic matter (biomass) and living species (biodiversity), and draws energy from the sun (renewable natural resource). The biosphere reveals an intrinsic tendency to self-organise and produce order (negentropy2 ). Over the centuries, when it found an efficient way to avoid transition to chaos thanks to internal self-control mechanisms, it actually settled as a system that works in cycles, without a tendency to increase or lessen, without wasting nonrenewable resources, and without polluting. This situation continued throughout several geological eras until the advent of the modern age. Man is also a part of the biodiversity of the system, an important but not essential part. While a biosphere without man is possible and existed in other geological eras, man without a biosphere is unthinkable. Even differentiating between man and biosphere is flawed, because man cannot be separated from his environment without losing his traits as a living being. Man contributes to the composition of the biosphere like a tree or bacteria. Unlike the balanced ecological system, industrial society has considered the natural environment as a receptacle from which it can extract everything it needs to 1 Ecofriendly or Eco-friendly adj. Having a beneficial effect on the environment or at least not caus-

ing environmental damage. (Collins English Dictionary—Complete and Unabridged 2012 Digital Edition ©HarperCollins Publishers). Origins: by 1993, from eco- + friendly. (Online Etymology Dictionary, ©2010 Douglas Harper). 2 The term “negative entropy” is the opposite of “entropy”, which in thermodynamics represents the index of the state of disorder of the physical system. It was introduced by Erwin Schrödinger in his book What is life? (1943). Léon Brillouin later summarised the concept using the word “negentropy” to express its intrinsic positive meaning.

1.1 Architecture/Ecology/Sustainable Development

3

survive and develop, and into which it can release the waste and rubbish of the processes that support it. Increasing anthropisation of the territory has compromised its functioning due to the disproportionate consumption of non-renewable resources, i.e., energy and matter. Several major effects of this anthropisation have sparked acknowledgement of the environmental problem faced by society. They include pollution of the air, water and soil, acoustic and electromagnetic pollution, climate change and its effect on man’s health, and hydro-geological changes. In particular, this awareness came about in two stages: the first was linked to the effects of local pollution, the knowledge that resources were depleting rapidly (a phase that started in the fifties and sixties), and the 1973 energy crisis. The second phase involved the concept of environmental protection, globally and from an intergenerational point of view. This prompted an assessment not only of the earth’s capacity to absorb these effects, but also the question of whether or not man’s activities were sustainable. Up until the mid-twentieth century the adjective sustainable (from the Latin sustin¯ere “to uphold”, sus- + -tin¯ere, combining form of ten¯ere to hold) was used to indicate “capable of being sustained” (Collins English Dictionary). Only later was its meaning extended to “able to continue over a period of time” (Cambridge dictionary). In the late eighties the definition in dictionaries had become “capable of being maintained at a steady level without exhausting natural resources or causing severe ecological damage: sustainable development” (Collins English Dictionary). Sustainability can be considered as the durability3 of complex systems, including living systems, capable of maintaining the possibility of adapting to the environment over a period of time (Bocco and Cavaglià 2008). The concept of sustainable development to which we refer today became a household word when it was used in the 1987 Report of the World Commission on Environment and Development (WCED) entitled “Our Common Future” in which the Chairman Gro Harlem Brundtland (at the time the Prime Minister of Norway) provided a definition enlarging the scope of the environmental topic to include associated social and economic effects. “Humanity has the ability to make development sustainable to ensure that it meets the needs of the present without compromising the ability of future generations to meet their own needs. The concept of sustainable development does imply limits—not absolute limits, but limitations imposed by the present state of technology and social organization on environmental resources and by the ability of the biosphere to absorb the effects of human activities (Brundtland 1987).” Although current results fall very short of the globally sustainable socio-economic development goals initially laid down by the World Summit, the studies and strategies on environmental protection, initially based on “short term” interventions, have led to the establishment of “long term” strategies. This is a decisive step towards sustainability. In fact, “sustainable development” involves the gradual growth of society (concept linked to the future) while respecting pre-existing social and ecological balances (concept linked to the present) or, in other words, “a will to ensure 3 In

French sustainable development is translated as développement durable.

4

1 Life Cycle and Sustainability: Concepts and Keywords

the development of society and wellbeing of individuals by using, as a limit, the capacity of the environment to sustain that development” (Lavagna 2008, p. 29). The concept of sustainability merges with the concept of development, juxtaposed against the concept of growth, characteristic of the neoclassical economic model. In fact, the current economic system is based on growth considered as a linear increase in the production of goods, to which the conditions of wellbeing are associated in proportion to the accumulation of goods and productivity. Wellbeing is measured by the Gross National Product (GNP) indicator. In the twentieth century the population increased 4 times, the GNP increased 14 times, industrial production 40 times (La Camera et al. 2003). These numbers are the result of the industrial revolution and its associated quantitative growth; on the one hand they prove that industrial production made it possible to increase the wellbeing of large swathes of society and, on the other, introduced an economic model based on continuous production involving continuous consumption (Lavagna 2008). The British economist David W. Pearce formulated an economic definition of sustainable development. According to Pearce “sustainable development involves maximising the net benefits of economic development, subject to maintaining the services and quality of natural resources over time or, in other words economic development is broadly construed to include not only increases in real pro capite incomes, but also other elements in social welfare” (Pearce and Turner 1990, p. 36). Two rules have to be respected in order to maintain the services and quality of natural resources over time: the use of renewable resources at a lesser or equal rate than that of the natural regeneration of the resource itself; optimisation of the use of nonrenewable resources, subject to sustainability between resources and technological progress. During that period Herman Daly, the American economist, father of the steady state economics theory4 and supporter of Ecological Economics,5 defined the principles of sustainable development for the management of resources as follows: • harvest rates should equal regeneration rate (sustainable yield) • waste emission rates should equal the natural assimilative capacities of the ecosystems into which the wastes are emitted. Daly believes that the relationship between the economy and ecology (which to a large extent still has to be established) must necessarily depend on sustainable balance (Tiezzi and Marchettini 1999). The intrinsic contradiction between economic and environmental sustainability is caused by the fact that a sustainable future on a planet with limited resources is irreconcilable with unlimited quantitative growth. Vice versa, if growth were coherent with the environmental system, then social policies would be at risk and enter into conflict with existing political dynamics. In

4 H. Daly, Steady State Economics, San Francisco, Freeman, 2nd ed. New York, Island Press, 1991. 5 The

economic theory developed by Professor Robert Costanza aimed at “developing sustainable models of economic development, unlike economic growth which is not sustainable in a finite planet” (Tiezzi and Marchettini 1999).

1.1 Architecture/Ecology/Sustainable Development

5

the future we must aim to achieve a balance between wellbeing-ecosystem-resource use. Designers should work “for a durable world even if not 100% sustainable!”6 To complete the general information about the concept of sustainable development it’s useful to cite the article published in May 1997 in the magazine Nature7 regarding the importance to invest in natural capital (main topic at the World Conference held in Stockholm in 1972).8 It was written by a group of researchers, including Robert Costanza, President of the International Society of Ecological Economics (ISEE). Scholars maintain “that services with which the environment contributes to the wealth of humanity are worth at least 33 trillion dollars per annum, while the Gross National Product (GNP) provided by human activities in one year is only 18 trillion dollars (Tiezzi and Marchettini 1999, p. 48). The study suggests a change in the accounts systems, bearing in mind the value of the services of the ecosystems and natural capital because, based on the above theory, wellbeing does not increase when GNP increases. Enzo Tiezzi maintains that “the new theories of sustainable development and ecological economics present us with a new paradigm: no longer an economy based on two parameters, labour and capital, but an ecological economy that acknowledges the existence of three parameters: labour, natural capital, and the capital produced by man” (Tiezzi and Marchettini 1999, p. 43).

1.2 Design Process/Multidisciplinary Approach/Life Cycle Approach One of the first times an “architect” took on the role of “site manager” was when Filippo Brunelleschi designed the dome of Santa Maria del Fiore (1418). Despite contemporary consolidated building techniques and the workmen’s expertise the latter would have been unable to build Brunelleschi’s project for which he also designed the worksite. Site management was the responsibility of the workmen, but from the Italian Renaissance onwards it passed to the designer who took on the new role of designer-builder.9 6 Lorenzo

Matteoli, Lesson 2018/03/04, Shanghai, China. Costanza et al., “The value of the world’s ecosystem services and natural capital” in Nature, 1997, 387, pp. 253–260. Reference taken from the book by Tiezzi E. and Marchettini N., Che cos’è lo sviluppo sostenibile, Donzelli Editore, Roma, 1999. 8 For the first time the Stockholm Conference in 1972 drew attention to the fact that to make longlasting improvements to people’s living conditions, natural resources had to be protected in a fair manner, and international collaboration was needed to achieve this goal. That same year the Club of Rome published the “The Limits to Growth” report which, against the background of the Stockholm Conference and the oil crisis in the early seventies, obtained worldwide resonance. The Stockholm Declaration, a crucial benchmark in international politics, anticipated the concept of “Sustainable Development” introduced in 1987. 9 In Italy the word progettista (designer) can refer to an architect, engineer, surveyor or appraiser. Elsewhere the word architect (English) or architecte (French) is used. 7 R.

6

1 Life Cycle and Sustainability: Concepts and Keywords

In the prologue of his manuscript De re aedificatoria (1485), Leon Battista Alberti considers “the architect, who by sure and wonderful reason and method, knows both how to devise through his own mind and energy, and to realize by construction, whatever can be most beautifully fitted out for the noble needs of man, by the movement of weights and the joining and massing of bodies. To do this he must have an understanding and knowledge of all the highest and most noble disciplines. This then is the architect.”10 The artisanal-style organisation that had characterised the building sector in previous centuries began to evolve when industrialisation and experimentation with new materials and innovative techniques were implemented between the late nineteenth and early twentieth centuries. New technologies required new building methods and better performance for components and materials which had to provide previously unattainable performance levels. Designer-builders were required to have more skills. These new experiments also involved the systems installed in buildings. Architects began to delegate the organisation and control of the environment to other specialists, a trend criticised by Reyner Banham who said it was above all society’s fault for not asking them to be something more than “simple creators of beautiful but inefficient inhabitable sculptures”. Banham believed that, unlike spontaneous construction, conscious architecture should provide unique, detailed solutions to specific problems (Banham 1969). Halfway through the twentieth century the concept of “quality management”11 was introduced into the way a design project is organised. In 1950 people began to realise that a design team must necessarily tackle an increasingly complex process that begins with the design and continues throughout the life cycle of the building or individual product. The goal is to reduce the construction period and maintain the same construction and planned maintenance costs so as to optimise investments and stay ahead of economic crises and inflation (Ciribini 1984). In architecture, especially when big investments are involved, efforts are made to apply the multidisciplinary Project Management approach based on the quality management concept and the time-costquality triangle or iron triangle (PMBOK Guide, 2008).12 In the design process, control is added to the design and execution stages; it involves experimentally ver-

10 Alberti

LB (1585), De re aedificatoria. On the art of building in ten books. (translated by Joseph Rykwert, Robert Tavernor and Neil Leach). Cambridge, Massachusetts: MIT Press, 1988. 11 The most important standard regarding quality is the ISO 9000 series (current edition 2015) Quality management systems—Fundamentals and vocabulary, initially issued in 2000. The standard is based on the first version of ISO 8402:1988 (Quality—Terminology). As regards quality in the building process, reference is made to standards: ISO 10006:2017 Quality management—Guidelines for quality management in projects and UNI 10722 (2007–2009) Edilizia - Qualificazione e verifica del progetto edilizio di nuove costruzioni. 12 In September 2016, the Ente Italiano di Normazione published standard UNI 11648 formally defining the figure of a project manager and the expertise and skills required to work professionally. Internationally, a project manager was defined by standard ISO 21500, later adopted and published in Italy as UNI ISO 21500 Guida alla gestione dei progetti (project management).

1.2 Design Process/Multidisciplinary Approach/Life Cycle Approach

7

ifying whether the required service, defined during the design stage, was actually achieved. The “designer” currently personifies a complex function executed by the project manager, or maître d’oeuvre in France who is responsible for coordinating the planning of the process, the drafting of the brief, the choice of the construction firm, and project management (Sinopoli 2004). Depending on how complex the project is, this function can involve more than one operator. The expression ‘participated design’ began to be used in the seventies. On the one hand it indicates a focus on the social client and, on the other, awareness of a “design platform” based on the analysis of collective needs. It interprets these needs and considers “the interrelated complexity of all the data”13 (Zanuso et al. 1977). In a debate on integrated design in 1977 Marco Zanuso encouraged “reappropriation” of the design process. He maintained that all the actors involved in a project—private and pubic clients, administrators, technicians and producers—must necessarily “assume and maintain in time and space the initiative for and responsibility of the project until it is completed, and beyond”.14 This means that roles, tasks and responsibilities have to be established during the early design stages so that all levels of the project can be successfully managed—design, execution, and use of the buildings—bearing in mind flexibility over time, associated social implications, and the overall cost of construction throughout the lifespan (Zanuso et al. 1977). The integrated design concept—associated here with a sort of economic and social ‘pre-sustainability’ (the term ‘sustainability’ was introduced ten years later in 1987)—began to also include the protection of the environment. In 1977 Gerard Blachère—Director of the Centre scientifique et technique du bâtiment (CSTB) from 1954 to 1974—declared that for the past twenty years design practice had already begun to consider the needs of the client and future users, the needs “of public authorities, for example energy saving or collective safety”, the needs “of the neighbours (e.g., noise) or local community (pollution)” and the requirements “regarding components and materials, as well as the need for a certain quality (i.e., performance)”.15 The performance-based approach began to be codified by the regulation authorities in the eighties,16 based amongst other things on Blachère’s principles (Blachère 1969). In the approach the ‘demand’ or ‘need’ is defined as “what is, by necessity, required for the correct implementation of an activity of the user or a technological function”.17 Where the ‘user’ represents the organisation, person, animal or object 13 M.

Zanuso (p. 1) in La progettazione integrata per l’edilizia industrializzata (series of general teaching debates)/M. Zanuso, N. Tubi, H. Weber [et al.], Milano: ITEC; 1977. 14 Ibidem. 15 Gerard Balchère, (p. 123) in La progettazione integrata per l’edilizia industrializzata (cycle of general teaching debates)/M. Zanuso, N. Tubi, H. Weber [et al.], Milano: ITEC; 1977. 16 In Italy standard UNI 8289:1981 classifies the requirements regarding safety, wellbeing, fruition, appearance, management, integration, and protection of the environment. In 1994 Maggi proposed two more standards: easy-assembly and economics. 17 UNI 10838:1999 Edilizia—Terminologia riferita all’utenza, alle prestazioni, al processo edilizio e alla qualità edilizia. This replaces the series UNI 7867 series dated 1978. (Internationally, ISO

8

1 Life Cycle and Sustainability: Concepts and Keywords

Fig. 1.1 The sustainable and quality approach, and a combination of the two (Thiébat 2013)

which uses, or is intended to use, a building or other construction works. It includes any person or entity who uses a facility, whether as occupant, visitor, member of the public, or other stakeholder.18 The requirements “involve turning a need into technical and scientific factors in order to identify how a building or its spatial and technical parts can satisfy said need, given certain conditions of use or stress”.19 Finally, the ‘performance’ “embodies the requirements according to univocally determined, assigned, or variable values defining the quality objective to be pursued by the project”.20 Sustainability and quality are the common goals of each of the specialist fields that are part of the design process. One senses that achieving these goals—as expressed by Banham, Zanuso, Blachère and many other scholars—is currently the ultimate aim of integrated design (Fig. 1.1). Satisfying the needs of users is the starting point on which to base and develop the organised sequence of the various stages of the building process; the latter includes the design process (planning, brief, definition of the metaprogetto, design stages) and the life cycle of the building itself (Bocco and Cavaglia 2008). In particular, the production and construction stages, use stage (operational, management, maintenance, functional and technological upgrade or recovery of the building), deconstruction and demolition of the building (including reuse and recycling of the materials and components), and regeneration of the soil (Sinopoli 2004).

19208:2016 Framework for specifying performance in buildings. It replaces several standards published in 1980 and in 1992–1994). 18 ISO 11863:2011 Buildings and building-related facilities—Functional and user requirements and performance. 19 UNI 10838:1999 op. cit. 20 Ibidem.

1.3 Time/Service Life/Future

9

1.3 Time/Service Life/Future The building process can be divided into three main phases (Fig. 1.2). The first, steered by the client and coordinated by the designer, is called “ideation”. This is when the design team is formed, the design brief is established, and the work is planned. The second phase can be called “construction”; during this phase the site manager and builder play leading roles. The third phase is the “service life” of the building. The user and manager of the building are the key players here, together with the community (indirectly involved). Despite the fact that many similarities exist between an industrial production process and a construction process, usually a building lasts much longer than any industrial product. This permanence, known as the “physical and functional service life” can be prolonged thanks to planned and unforeseen maintenance21 to upgrade the functional and technological features of the building. This necessarily means that the design team must envisage the work that has to be planned and provide the final user with a maintenance plan (be it the client, end-user or manager). In order to make this activity operational and useful it should be drafted based on an analysis that takes into account functional, environmental and economic needs and requirements, bearing in mind a Life Cycle Thinking that considers a pre-established period of analysis. Such analysis should include, for example, Life Cycle Assessment and Life Cycle Costing (illustrated in detail in the third chapter) which can be integrated into the building process as illustrated in chapter four (i.e., Life Cycle Design Model). To make the building a future proof system the designer must anticipate reduction in performance over time; during the design stage he must also provide a solution for the requirements that will have an effect in the future. The future-proof concept applied to the field of construction represents the relationship between the use stage and the design stage to satisfy the goals of sustainability and resilience throughout the life of the asset (Pitts 2008; Georgiadou et al. 2012; Rehman and Ryan 2018). On this issue European regulatory authorities are promoting long term refurbishment strategies and planning actions and measures to enhance sustainability and energy efficiency in the building life cycle. For example, the new direc-

Fig. 1.2 The diagram shows the building process divided into three main phases (ideation, construction, and service life) and the stakeholders involved in the life cycle 21 For

example the work that has to be performed after accidental damage or technological obsolescence.

10

1 Life Cycle and Sustainability: Concepts and Keywords

tive 2018/844/EU modifying the directives on energy performance in buildings (2010/31/EU) and energy efficiency (2012/27/EU) suggests that European States “could make use of concepts such as trigger points, namely opportune moments in the life cycle of a building, for example from a cost-effectiveness or disruption perspective, for carrying out energy efficiency renovations”.

1.4 Regulations/Standards/Codes The relationship between sustainability and life cycle is becoming a key issue for designers, property developers, final users, citizens, and public administrations. Developers concentrate more on the global costs they have to bear during the economic life of the investments. Instead final users focus more on the environmental and economic impact of the products (including buildings), especially if the operational and disposal stages are taken into consideration. Public authorities have to provide concrete answers to the growing awareness of local communities regarding sustainability (Ryan 2014). In this context the non-governmental agencies voluntarily involved with drafting standards, for instance the International Organization for Standardization (ISO), the European Committee for Standardization (CEN) and the Ente Nazionale Italiano di Unificazione (UNI), are developing methods and tools to extend the use of environmentally, economically, and socially compatible practices. In the world of construction committees have begun to warm to the idea of sustainability, both internationally—i.e., the activities of the ISO-TC59-SC17 (sustainability in building constructions) and the CEN TC350—and in Italy, thanks to the work of the UNI/CT033/GL2 (Sustainability in the building industry). This has led to the gradual implementation of regional laws and provincial and municipal standards. In particular, several technical commissions and working groups have jointly tackled the issue of sustainability vis-à-vis the life cycle and prompted the definition of the Life Cycle Design Model presented in this book: • ISO-TC59 (SC14 and SC17)—Buildings and civil engineering works. • CEN TC 350—Sustainability of construction works.

1.4.1 ISO/TC 59 Buildings and Civil Engineering Works (SC14 and SC17) The International Organization for Standardization (ISO) is a global federation of national standards bodies involved in developing standards. Each interested body can take part in the technical commissions proposed by the ISO. Figure 1.3 presents several working groups of the ISO-TC59 Buildings and Civil Engineering Works, in particular the subcommittees: SC14—Design Life, and SC17—Sustainability in

1.4 Regulations/Standards/Codes

11

Fig. 1.3 ISO-TC59—Buildings and civil engineering works (SC14 and SC17) showing some of the standards used to develop the LCDM presented in this book

building constructions. Subcommittee 14 developed the service life standards, including Standard 15686-5 regarding life cycle costing.22 The key goal of the second group, SC17, is to develop standards to harmonize terminology and methodologies with reference to the three fields of sustainability for buildings and components. For example, WG4 of the SC17 has developed standards regarding the Environmental performance of buildings, i.e. ISO 21929-1 (current version dated 2011) Sustainability indicators—Part 1: Framework for the development of indicators and a core set of indicators for buildings and ISO/TS 21931 (current version dated 2010)—Framework for methods of assessment of environmental performance of construction works. The former is methodological. It describes and provides guidelines for the development of sustainability indicators related to buildings and defines the aspects of buildings that need to be considered when developing systems of sustainability indicators. The second standard provides a general reference framework for improving the quality and comparability of methods to assess the environmental performance of buildings. It identifies and describes issues to be taken into account when using methods to assess the environmental performance of new or existing building properties in the design, construction, operation, refurbishment, and deconstruction stages.23 ISO 21931-1 aims to bridge the gap between regional and national methods by providing a common framework. It applies to all stages of a construction project, from design through to construction, operation, maintenance, refurbishment and deconstruction, to ensure that the finished product is an eco-friendly building. It is intended be used in conjunction and in line with the principles set out in the TC 207 that includes the ISO 14020 family of International Standards on environmental 22 Cfr.

Chapter three.

23 https://www.iso.org/standard/45559.html?browse=tc

(access 24/09/2018).

12

1 Life Cycle and Sustainability: Concepts and Keywords

Fig. 1.4 ISO-TC59-SC17. Working group on the environmental performance of buildings

labelling, and also ISO 14040 on Life Cycle Assessment (Fig. 1.4) and ISO 15392 on general principles of sustainability in building construction.24

1.4.2 CEN/TC 350 Sustainability of Construction Works The Conference organised in 2001 by the European Producers of Materials for Construction (CEPMC) highlighted the need for this sector to define harmonised standards for environmental brands, especially for Environmental Product Declarations (EPD). In 2004 a European Technical Committee was set up, i.e., the CEN/TC 350 “Sustainability of Construction Works”. Its brief was to introduce and ensure the application of tools and methods for the environmental, economic, and social sustainability of new or existing buildings, in other words the three pillars of sustainability introduced by Agenda 21. The Committee is currently working synergistically with existing international standards, including the ones developed by the aforementioned groups (ISO-TC59-SC17 and ISO-TC59-SC14). The members of the Task Groups and Working Groups are representatives of the standardisation authorities of individual Member States (UNI, BSI, AFNOR, etc.). The latter can nominate experts and delegates (e.g., universities, producers of construction materials, etc.) to develop the standards proposed by the Technical Commission.

24 https://www.iso.org/news/2010/08/Ref1344.html

(access 24/09/2018).

1.4 Regulations/Standards/Codes

13

Fig. 1.5 Standards developed by CEN/TC 350 “sustainability of construction works”

Like the ISO structure, there are seven working groups in the CEN TC/350. These groups focus on specific issues such as: Environmental performance of buildings; Product Level (EPD’s, communication formats etc.); Economic Performance Assessment of Buildings; Social Performance Assessment of Buildings; Civil Engineering works; Framework and coordination; Sustainable refurbishment. Apart from the topic-oriented working groups, the TG “framework” was set up to describe the tools and general principles for the assessment of environmental, economic and social sustainability (Fig. 1.5). The national groups that are currently focusing on the same issues, for example the Italian group UNI/033/CT02 Sustainability in the building industry, have transposed the CEN standards after working jointly with other countries. This has led to harmonised European methodologies especially as regards the stages of the entire life cycle of buildings and the adoption of the broadest concept of sustainability, in other words one which considers environmental and economic aspects and the interconnected social implications. The basic principle underlying the standards is in line with EU sustainability policies as well as the legal frameworks related to those policies, such as: • DIRECTIVE 2008/98/EC on Waste. • DIRECTIVE 2009/125/EC Establishing a framework for the setting of ecodesign requirements for energy-related products. • DIRECTIVE 2010/31/EU on the energy performance of buildings. • DIRECTIVE 2012/27/EU on energy efficiency. The EU has set itself a 20% energy savings target by 2020 (compared to the projected use of energy in 2010). In 2016 the Commission proposed an update to the Energy Efficiency Directive including a new 30% energy efficiency target for 2030.25 25 On

14 June 2018 the Commission, the Parliament and the Council reached a political agreement including a binding energy efficiency target of 32.5% for the EU in 2030, with a clause for an

14

1 Life Cycle and Sustainability: Concepts and Keywords

The 2010 Energy Performance of Buildings Directive and the 2012 Energy Efficiency Directive are the EU’s main legislative instruments promoting the improvement of the energy performance of buildings within the EU. Following the introduction of energy efficiency requirements in national building codes in line with the Directive, new buildings currently consume only half as much as typical buildings constructed in the eighties. Directive 2018/844/EU26 published on 19 June 2018 amended the Energy Performance of Buildings Directive. This revision introduced targeted amendments to the current Directive to accelerate the cost-effective renovation of existing buildings, with a view to achieve a decarbonised building stock by 2050 and mobilise investments.27 The revision also supports electromobility infrastructure deployment in the car parks of buildings and introduces new provisions to enhance smart technologies and technical building systems, including automation. It is therefore possible to state that in the last two decades the European standards system has helped to increase the number of energy efficient buildings during use (both new and existing). It has, however, not been as incisive in lowering the energy quota linked to production and implementation of building products. In fact, lowering the energy requirement of buildings during use increases the importance of not only the supply of the materials, but also the production and installation of the building products and components used in the construction of buildings (Giordano et al. 2017).

1.5 Integrated Design/Assessment Systems/Tools Considering a building over its entire life cycle means taking into equal account functional, emotional, material, and social levels (Drexler and El Kohuli 2012). Architecture must be user-oriented, usable, and adaptable to mutable needs. It also has to be durable and strong enough to tackle economical, societal and climate changes. It must also be aware of all the energy and material flows that occur throughout the life of the building. ‘Housing quality’ is a multilayered construct of objective factors and individual needs and values. “If architecture is to fulfil an integrated expectation, then sustainable building cannot be reduced to quantifiable and measurable aspects but must be considered with the broad spectrum of human needs” (Drexler and El Kohuli 2012, p. 58). An architect’s creative act must be extended to include a systemic approach to design, combining creativity with an analytical and scientific process. However this goal must be shared by all design team members. Sustainability assessment methods began to be developed in the nineties as part of a holistic design (Mendler and Odell upwards revision by 2023. This political agreement must now be formally adopted by the European Parliament and the Council (https://ec.europa.eu/energy/en/topics/energy-efficiency). 26 http://data.europa.eu/eli/dir/2018/844/oj. 27 Member States will have 20 months to transpose its provisions into national law (namely by 10 March 2020).

1.5 Integrated Design/Assessment Systems/Tools

15

2000) in which every performance requirement could be assessed individually, but also as part of a more complex system that considered the building as the merger of architecture, structure, systems, and outdoor areas. The latest sustainability certification systems not only use indicators and points to formalise the level of balance between the various factors impacting on achieving sustainability (energy, environmental, social, and economic factors), but also try to take into consideration the entire life cycle of the building (Attia et al. 2012; König et al. 2010). An ‘informed’ use of sustainability assessment methods provides the architect coordinating the team of specialists and less experienced users with a model, a guideline he can use very early on in the design process to influence the design choices and share them with the actors involved. Rating systems force design teams to think holistic and, at the same time, reveal which sustainable objectives must be improved (Østergård et al. 2016). Unfortunately the ‘uninformed’ use of these tools, sometimes too complex and under-exploited during the early stages of the design process, leads to a mere collection of indicators hampering the potential achievement of the sustainability goals set by these protocols. Often this is caused by a number of factors: the fact that performing these analyses is either not remunerated or remunerated very little; modest integration between expert sustainability assessment professionals in the design team and commissioning agencies; frequent discrepancy between the standard design system and protocol standards (often due to temporal or territorial inconsistencies). Furthermore, unlike energy certificates, in today’s world these certificates are always voluntary tools that hardly ever provide access to volumetric or economic incentives. The following are some of the environmental protocols used in Europe and in Italy: DGNB (DE), LEED (USA + EU), Protocollo ITACA (PdR UNI/13:2016) (IT) and BREAM (UK). They are defined “rating systems” because every performance indicator is linked to a score. When the total is normalised based on a weighting system that depends on the intended use of the building it provides the building’s degree of sustainability.28 Indicators can be qualitative or quantitative. They are grouped in impact categories and can be considered in a more or less detailed manner (Fig. 1.6). Analytical methods providing complex results, such as Life Cycle Assessment and Life Cycle Costing, can be used to calculate environmental and economic indicators if the assessment system envisages an analysis that includes the entire life cycle (e.g., DGNB). In the last ten years many regulatory contributions in Europe have tried to standardise and harmonise assessment methods in order to achieve sustainable construction. In addition designers and users have inputted into the efforts to increase eco-compatible and energy efficient design choices. Nevertheless, the fact these assessment systems are complex and difficult to integrate into design practices in terms of time and costs has prevented widespread and large-scale application of these systems to “normal” 28 For

more in-depth information about sustainability certification systems, consult the websites of the individual assessment systems. The continuous upgrading of these systems and the close link between geography/standards and the application of the system requires accurate verification that has to be performed on a case by case basis according to the use requirements.

16

1 Life Cycle and Sustainability: Concepts and Keywords

Fig. 1.6 In the figure comparing several assessment systems, dark grey indicates an in-depth level of assessment of a category of criteria; light grey indicates a category with a lower level of assessment; white indicates that the category has not been considered

buildings (i.e., small private and public buildings or buildings with reduced media impact). In fact they have been used only for buildings which have required big and mainly private investments. This consideration inspired the decision to develop a method based on just a few performance indicators and then integrate it into design practice, as proposed in this book (cfr. Chap. 4). Several studies emphasise the following: when concise, life cycle-based indicators (e.g., environmental impact indicators) are used and applied during design, and combined with the energy certification of the building, this would ensure the dissemination of more sustainable design solutions. Furthermore, the “transparent” use of individual indicators would avoid the risk of a prescriptive approach using standard design solutions (assessment systems define performance levels) and replace it with a performance-based approach capable of stimulating innovation and continuous improvement (Lavagna 2008).

1.5.1 Tools for Integrated Design Integrated design needs a holistic approach to manage multiple factors; as a result, the interchangeability of design tools becomes a key issue. Building Information Modelling (BIM) defines the “use of a shared digital representation of a built object (including buildings, bridges, roads, process plants, etc.) to facilitate design, construction and operation processes to form a reliable basis

1.5 Integrated Design/Assessment Systems/Tools

17

for decisions” (ISO 29481-1:2016). The US National Building Information Model Standard Project Committee has developed the following definition: “Building Information Modelling (BIM) is a digital representation of the physical and functional characteristics of a facility. BIM is a shared knowledge resource providing information about a facility and forming a reliable basis for decisions during its life-cycle, defined as existing from earliest conception to demolition”.29 BIM is intended to facilitate interoperability between software applications used during all stages of the life cycle of construction works, including briefing, design, documentation, construction, operation, maintenance, and demolition. It promotes digital collaboration between actors in the construction process and provides a basis for accurate, reliable, repeatable and high-quality information exchange. The European Union Public Procurement Directive (EUPPD) of 15/01/2014 recommended the use of electronic instruments supported by BIM technology, acknowledging the possible advantages they provide. In Italy the Directive was transposed by the Procurement Code (Decree Law 50/2016) and subsequent implementation decree.30 Using BIM makes it possible to keep track of the design data and decisions over time, ensuring immediate availability of the maintenance plan and the as-built model, always updated and with all the documentation. BIM methodology envisages that the design model (BIM) evolves and becomes a building model (Project Information Model—PIM) and then a management model (Asset Information Model—AIM) throughout the commission, complete with information and useful data for the users who will have to manage and maintain the building. Merging design parameters quantifying the environmental aspects (e.g. LCA) and life cycle costs (e.g., LCC) could be facilitated not only by the widespread use of BIM, but also by the integration of these analyses with the rating systems. In fact, in the last ten years software companies have joined forces and created several partnerships to merge environmental impact calculation tools and rating systems with 3D modelling software. The latter are generally data exchange interfaces with databases and calculation methods developed and continually updated by public and private research agencies. LCA software is based on national databases and internationally acknowledged impact analysis methods (e.g., the IPCC method to quantify the Global Warming Potential). The most complete and sophisticated LCA software31 are usually used at the end of the design process, or even during construction, when design choices are final and the established quantity of the materials makes it almost impossible to compare alternative solutions (Dalla Valle et al. 2016). 29 http://www.nationalbimstandard.org/faqs

(access 1/10/2018). use of BIM in public tenders in Italy will be mandatory on 01/01/2019 for complex works (over 100 million); 01/01/2020 for works between 50 and 100 million; 01/01/2025 for works worth one million euro. 31 These software contain databases and impact assessment methods. They provide access to the data in the databases and calculation methods, ensuring traceability, transparency and personalised data. These software programmes can be used only by expert users familiar with LCA. The most popular in Europe include SimaPro (Pré consultant, NL); GaBi (DE) and the OpenLCA (DE). 30 The

18

1 Life Cycle and Sustainability: Concepts and Keywords

Fig. 1.7 The building life cycle stages shown in Fig. 1.2 and more in-depth data quality

The widespread use of BIM has however helped to develop simplified tools that calculate environmental impacts (e.g., LCA)32 and other aspects of sustainability (e.g., rating systems) based on the data in 3D models. By doing so they facilitate assessment and make the results more comprehensible to less expert users. This makes it possible to consider environmental impact and other factors given the increasing amount of available data throughout the design process (Thiébat and Cocina 2018; Thiébat 2013). Using integrated software enables parametric quantitative data to be considered during the early design stages (Cheung et al. 2012), followed by secondary data from the databases (or price lists in the case of costs), and finally primary data based on technical sheets, transportation documents (DDT), and the EPDs33 of specific products (Fig. 1.7). The model can be updated at the end of the construction process by inserting the products and exact quantities. During use it can act as a “receptacle” to contain management data used to assess the impact on the life cycle and optimise use, i.e., save environmental and economic resources. Several researchers have focused on improving the interoperability between BIM and specialised tools for LCA and the energy analysis of buildings (Moon et al. 2015). However this approach—using a chain of specialised software tools—increases complexity and error transfer, thus reducing adoption by building designers and stakeholders. As a result, a framework/prototype is still needed to achieve comprehensive integration of LCA into a BIM tool in order to enable designers to easily 32 Although these software are based on databases and assessment methods acknowledged by the scientific community, they have a simplified interface where data cannot be modified. The validity of the results is therefore compromised by factors linked to data localisation and insufficient transparency. These software can be used by non-expert users unfamiliar with LCA and are very useful during the initial stages of a project to compare different building solutions. They include, amongst others: eToolLCD (AUS), Tally (India), Bees (USA), IESVE (UK) and OneClickLCA (Finland). Some of them can import the quantities and characteristics of materials directly from a BIM model into a 3D modelling software, including Revit (Autodesk, USA), Archicad (Graphisoft, Ungheria). 33 The Environmental Product Declaration (EPD) is created and verified in accordance with ISO 14025. It is based on the Life Cycle Assessment specified in ISO 14040 and ISO 14044.

1.5 Integrated Design/Assessment Systems/Tools

19

compare different design alternatives during the early design stage (Nwodo et al. 2017). These topics are part of the work of the JRC (Joint Research Centre), the European Commission’s science and knowledge service. The JRC carries out research and provides independent scientific advice in support of EU policies. In 2017 the JRC worked closely with industry stakeholders to develop a common framework called Level(s): a set of simple metrics used to measure the sustainability performance of buildings throughout their life cycle. It is based on existing tools and standards, and covers all the areas of energy, materials, water, health, comfort, climate change, and life cycle cost and value. Level(s) provides a common “sustainable” language for sustainability tools and certification schemes. The initiative is open source and freely available (Dodd et al. 2017).34

References Alberti LB (1585) De re aedificatoria. On the art of building in ten books. (translated by Joseph Rykwert, Robert Tavernor and Neil Leach). Cambridge, Massachusetts: MIT Press, 1988 Attia S, Gratia E, De Herde A, Hensen JLM (2012) Simulation-based decision support tool for early stages of zero-energy building design. Energy Build 49:2–15 Banham R (1969) The architecture of the well-tempered environment. Arch Press, Londra Battaglia A (a cura di) (2006) Saint-John Perse, Uccelli, Ed. dell’Orso, Alessandria Blachère G (1969) Savoir batir. Eyrolles, Paris Bocco A, Cavaglià G (2008) Cultura tecnologica dell’architettura. Pensieri e parole, prima dei disegni, Carocci, Roma Brundtland GH (1987) Our common future: report of the world commission on environment and development. Oxford University Press Cheung FKT, Rihan J, Tah J, Duce D, Kurul E (2012) Early stage multi-level cost estimation for schematic BIM models. Autom Constr 27:67–77 Ciribini G (1984) Tecnologia e progetto. Celid, Torino Dalla Valle A, Lavagna M, Campioli A (2016) Strumenti LCA di supporto al settore delle costruzioni in X Convegno dell’Associazione Rete Italiana LCA 2016 Life Cycle Thinking, sostenibilità ed economia circolare Ravenna 23–24 giugno 2016 Dodd N, Cordella M, Traverso M, Donatello S (2017) Level(s)—a common EU framework of core sustainability indicators for office and residential buildings, Part 1, 2 and 3. European Commission, JRC Drexler H, El Kohuli S (2012) Holistic housing. Concepts, design strategies and processes, DETAIL special, Munich, Germany Gauzin-Muller D (2012) A short history of sustainable architecture. In: Drexler H, El Kohuli S (eds) Holistic housing. Concepts, design strategies and processes, DETAIL special, Munich, Germany Georgiadou MC, Hacking T, Guthrie P (2012) A conceptual framework for future-proofing the energy performance of buildings. J Energy Policy 47:145–155 Giacchetta A, Magliocco A (2007) Progettazione sostenibile. Carocci editore, Roma Giordano R, Serra V, Demaria E, Duzel A (2017) Embodied energy versus operational energy in a nearly zero energy building case study. Energy Procedia 111:367–376 König H, Kohler N, Kreissig J, Lützkendorf T (2010) A life cycle approach to buildings. In: principles, calculations, design tools. Detail Green Books, Regensburg 34 http://ec.europa.eu/environment/eussd/buildings.htm

(access on 31/10/2018).

20

1 Life Cycle and Sustainability: Concepts and Keywords

La Camera F (2003) Sviluppo sostenibile. Origini, teoria e pratica, Editori Riuniti, Roma Lavagna M (2008) Life cycle assessment in Edilizia, Hoepli, Milano Manfron V, Siviero E (1998) Manutenzione delle costruzioni: progetto e gestione. UTET, Torino Mendler SF, Odell W (2000) The HOK guidebook to sustainable design. Wiley, New York Moon H, Kim H, Kamat VR, Kang L (2015) BIM-based construction scheduling method using optimization theory for reducing activity overlaps. J Comput Civil Eng 29(3), Kim and Anderson 2012 Nwodo MN, Anumba CJ, Asadi S (2017) BIM-based life cycle assessment and costing of buildings: current trends and opportunities. In: ASCE International Workshop on Computing in Civil Engineering Østergård T, Jensen RL, Maagaard S (2016) Building simulations supporting decision making in early design: A review. Renew Sustain Energy Rev 61(August):187–201 Pearce DW, Turner RK (1990) Economics of natural resources and the environment. The Johns Hopkins University Press, Baltimore Pitts A (2008) Future proof construction—future building and systems design for energy and fuel flexibility. J Energy Policy 36:4539–4543 PMBOK (2013) A guide to the project management body of knowledge (PMBOK Guide), 5th edn. Project-Management-Institute, Upper Darby, PA Rehman OU, Ryan MJ (2018) A framework for design for sustainable future-proofing. J Clean Prod 170:715–726 Ryan MJ (2014) Design for system retirement. J Clean Prod 70:203–210 Sinopoli N (2004) La tecnologia invisibile. Il processo di produzione dell’architettura e le sue regie, Angeli, Milano Steele J (2005) Ecological architecture—a critical history. Thames and Hudson, NY Terzi C (2001) I Piani Della Luce, Domus, Milano Thiébat F (2013) Life-cycle design for sustainable architecture. Techne J Technol Archit Environ 05:177–183 Thiébat F, Cocina G (2018) The multidisciplinary approach for life cycle architecture. In: Abstract Book, 24th ISDRS Conference, Action for a Sustainable World: From Theory to Practice, Messina, Italy, 13–15 June 2018 Tiezzi E, Marchettini N (1999) Che cos’è lo sviluppo sostenibile. Donzelli Editore, Roma Zanuso M, Tubi N, Weber H (ed) (1977), La progettazione integrata per l’edilizia industrializzata (cycle of general teaching debates), ITEC, Milano

Chapter 2

The Environment and Economics

Abstract This chapter outlines the relationship between environmental and economic aspects by briefly describing the main concepts. All the topics presented here represent a framework that can be explored more in-depth using the references provided. They will help the reader find the fil rouge of a sustainable approach linking economics and environmental architecture.

2.1 The Circular Economy Until recently economic theory neglected the natural environment as an economic resource; natural resources such as the air, the seas, the carrying capacity of the environment, etc., were considered to be inexhaustible. By definition, the economy focuses only on assets and resources that are scarce, i.e., available in limited quantities compared to the needs to be met. The problem of satisfying needs wouldn’t exist if we had unlimited resources. However, since resources, including environmental resources, are scarce and limited, we must establish criteria to distribute and use them efficiently in order to ensure collective wellbeing. One of the objectives of the economic system is to divide societal resources between alternative uses. The environmental economics theory illustrates how an economic system can efficiently use resources, and identifies possible causes (Panella 2002). The kernels of a new approach that was later to be adopted were presented in several scientific contributions written in the sixties; the approach inspired the elaboration of a new economic theory called Environmental Economics (Ghisellini et al. 2016). Several international studies emphasised the fact that economic development triggers considerable costs, including environmental costs. The environment is an economic resource; it is used to satisfy the needs created by human activities either directly (air to breathe, etc.) or indirectly when resources are used in production processes. Figure 2.1 illustrates the interactions between the economic and environmental systems. The upper area represents the economy and the inter-economic linkages that occur in it (e.g., supply and demand), while the lower area represents the envi© Springer Nature Switzerland AG 2019 F. Thiébat, Life Cycle Design, PoliTO Springer Series, https://doi.org/10.1007/978-3-030-11497-8_2

21

22

2 The Environment and Economics

Fig. 2.1 Interaction between the environment and the economic system (re-elaboration of the image, Pearce 1990)

ronment. The latter includes all in situ resources—energy sources, fisheries, land, and the capacity of the environment to assimilate waste products. The producers of primary products that supply producers of final products (e.g., how the production of cars affects the demand for steel) is an example of “inter-economic linkages”. How forests affect water supply and soil quality is an example of “inter-environmental linkages”. Here there is no economic dimension (no exchange value). The British economist David W. Pearce (1990) highlights how, in primis, the entities within the environmental matrix (the square at the bottom) appear not to be part of the economic dimension. Environmental Economics is concerned with both areas, as well with the interactions between them (e.g., how the demand for steel affects the demand for water). Environmental Economics thus tends to be more holistic than conventional economics. These interactions are established by the flow of environmental assets and natural resources (as defined below) inputted into the economic system by the environment (in Fig. 2.1: input into the economic system by the environment) and by the flow between the economic system and the environment (waste discharge) pursuant to consumption (in Fig. 2.1: flow of waste from the economic system to the environment). Environmental resources and assets (air, forests, water, etc.) and natural resources (coal, oil, zinc, etc.) can either be reproduced or renewed; however, in some cases they are non-renewable. Renewable resources have flow and stock problems, while non-renewable resources only have problems of stock and optimisation of their exploitation (Table 2.1). The problem linked to non-renewable resources is temporal and involves establishing optimal exploitation of a resource stock. Instead renewable resources generate two problems. The optimal exploitation of the resource flow has to be calculated and, at the same time, ensure that the stock does not drop below a certain level in order

2.1 The Circular Economy

23

Table 2.1 Table of natural and environmental resources Type of resource

Characteristics

Example

Unlimited renewable resources

Unlimited for the entire life of the human race

Solar energy, wind energy, waves and tides

Limited renewable resources

Limited depending on the rate of renewabilitya

The atmosphere, carrying capacity of the environment, water, land, etc

Recyclable non-renewable resources

The stock can vary, but services can be recycled

Metals (copper, zinc, iron, …)

Non-renewable resources

Establishing the optimal use rate

Oil

a The exploitation rate of the resource must be lower or equal to the rate of replacement and regrowth

of the resource itself

not to compromise natural reconstitution. In other words, the exploitation rate of the resource must be lower, or equal, to the rate of replacement and regrowth of the resources itself.

2.1.1 The Carrying Capacity of the Environment The first law of thermodynamics, according to which it is not possible to create or destroy energy and matter, establishes the relationship between the resources removed from the environment and the total amount of waste. Everything we take from the environment cannot be destroyed but only converted and dispersed. In light of these considerations we can imagine turning the linear economic system into a circular system1 in which recycling can convert part of the waste into a resource. Anything that is not recycled is discarded into the environment. The second law of thermodynamics states that in every transformation of energy, part of the energy is dispersed in a form that cannot be used to perform more work. According to the economist Nicholas Georgescu-Roegen, the second law provides an answer to the fact that it is not always possible to recycle and that, sometimes, this is not a lost opportunity. Recycling may in fact not be possible within an economic system because the raw materials that are used often tend to be dispersed in the system itself, or else because some resource categories (e.g., energy resources) cannot be recycled. Entropy, the measurement of energy no longer available, is a physical obstacle to the economic-environmental system. The environment is able to assimilate waste and convert it into less dangerous and ecologically useful substances: this is called carrying capacity. So long as the quantity and quality of discharged waste is proportional to the carrying capacity of the environment the circular economic system will function as a natural system. If instead 1 For

more clarification, review the theory of the circular system in Pearce and Turner (1990).

24

2 The Environment and Economics

the discharged waste damages the carrying capacity of the ecosystem, the economic function of the environment as a receptor of waste will be damaged, transforming what could have been a renewable resource into a non-renewable resource (Pearce and Turner 1990). Table 2.1 shows that “the carrying capacity of the environment” is deemed to be a finite renewable resource, due to its renewable rate, considering that the exploitation rate of the resource must be lower or equal to the rate of replacement and regrowth of the resource itself. The materials balance model schematically illustrated in Fig. 2.1 identifies the economic functions2 of the natural environment. With this in mind, Pearce and Turner define the environment as a system which: • generates resources as input into the production system • has a carrying capacity: capacity to assimilate waste and convert it into less dangerous and ecologically useful substances • is directly useful in the form of aesthetic enjoyment of the landscape and spiritual tranquillity. When exploitation of the environment goes beyond its carrying capacity, then the direction in which we are moving is no longer one of preservation.

2.2 Environmental Economics Economics is a discipline that studies how to allocate resources to produce goods and services in order to satisfy needs. Scarce resources, i.e., the fact they are useful and available in limited quantities compared to the demand, is the reason why we talk of economic resources. Up until the seventies and, in particular, the 1973 energy crisis, energy produced by fossil, liquid and gaseous fuels seemed almost unlimited. That was, however, the moment we began to realise that resources were finite. Traditional economics examines how to combine (scarce) resources and (abundant) needs. In this case “needs” should be considered as the ensemble of goods and services that are bought/sold on the market, while “resources” are the sources required to produce these goods and services; they can be natural (the earth), human (labour) or produced (capital). The goal of the economy is therefore to produce goods and services. In classical economics the relationship between the quantity of goods exchanged on the market and its price is the basic tenet of the supply and demand theory according to which consumers’ preferences determine the demand for goods, while the costs sustained by companies establish their supply. Market equilibrium is established by price and quantity thus creating balance between supply and demand. In Fig. 2.2, market price p* is the price corresponding 2 They

1990).

are economic functions because they all have a positive economic value (Pearce and Turner

2.2 Environmental Economics

25

Fig. 2.2 Market equilibrium

to the point where supply and demand meet at a certain moment in time, while q* is the point when the quantity of foodstuffs consumers wish to purchase is the same quantity the producers want to sell (point of market equilibrium). In a classical economic model, waste has no “value”3 and is therefore not considered. As mentioned earlier, the classical economic model affects the environmental system, especially as regards two specific topics: 1. Non-renewable environmental resources: the market manages scarce resources in the best way possible, but does not have the necessary tools to stop exhaustion of resources. Environmental resources are in fact limited and therefore exhaustible. 2. Pollution: factors such as water, air, and space, involved in the production system, cannot be managed by the market economy if they have no (exchange) value. When environmental problems burst onto the scene it became apparent that using the tools of classical theory to manage them would be difficult if not impossible since natural resources were considered as unlimited or devoid of an exchange value; as such they could not be inserted in the self-regulating mechanism of the market. To tackle these problems the environmental economy appeared in the early seventies as a branch of classical economics. It focused primarily on pollution problems and proposed to deal with the problems of the environment by integrating the external effects of economic activities into the market: internalisation of externalities. This means assigning a price to the “free” goods (air, water, etc.) that had artificially become “scarce” and taking them into account when calculating production costs. Two opposing positions emerged regarding resource exhaustion: • “We have to worry about the danger of exhaustion”. A pessimistic view, also called “The Economics of the Coming Spaceship Earth”. • “Technology will always compensate exhaustion”. Optimistic view, also called “Frontier Economics”.

3 Considered

as an exchange value on the market.

26

2 The Environment and Economics

The 1972 Limits to Growth report that MIT commissioned the Club of Rome (Meadow et al. 1972a, b) established that the limits to development on this planet will be reached at an unspecified moment within the next 100 years if nothing is done to curb the current growth rate of the population, industrialisation, pollution, production of food, and exploitation of resources. The result will probably be a sudden and uncontrollable decline in the population and industrial output. In the same years McKelvey’s study on the classification of world mineral resources/reserves (Falkie and McKelvey 1976) for the United States Bureau of Mines (USBM) highlighted the fact that known and economically available resources could represent only one part of existing resources. The effects of the exhaustion warning4 on the system can be deemed to be positive; they triggered the following acknowledgements: • Technology permits replacement, but does not stop resource exhaustion; • If we need to be concerned about exhaustion, traditional market mechanisms will not stop it. Environmental pollution is caused by contamination (discharge, waste, etc.) from human activities. Nevertheless, the economic definition of pollution includes the physical effect of waste on the environment and man’s reaction to that physical effect which can be biological (e.g., human health), chemical (e.g., climate change), or acoustic (e.g., noise). Man’s reaction is tangible in “loss of wellbeing” and therefore “usefulness”. Economists consider pollution an external cost5 that occurs only if one or more individuals loose that wellbeing (Pearce and Turner 1990). The main problem associated with negative externalities is the fact that those who produce the damage do not have to bear the whole cost of that damage: this causes “market failure”, i.e., the market can no longer guarantee the social efficiency of resource allocation. The internalisation of externalities is a possible solution (Musu 2000). Environmental economics has used the internalisation of externalities principle to tackle and in many ways attempt to solve the problems of pollution, but there are enormous difficulties, for example the identification/awareness of the damage, quantification of the damage, and how to remove it. Being aware of the damage is often complicated because we ignore the kind of damage caused by many new products. There are numerous examples of new products that did produce damage, but that damage was acknowledged only much later (e.g. DDT, CFFC, asbestos). It’s even more difficult to quantify damage, because it repeatedly affects free goods without a market price. As far as damage removal is concerned, if the damage is slight, then one option may be to reimburse the damaged party; if instead severe damage has been caused, actions must be taken to minimise it. Economists do not usually suggest eliminating externalities because they maintain that the optimal level of externalities is not equal to zero. From an economic point of 4 Cfr.

Meadows et al. (1972a, b), (Falkie and McKelvey 1976), Boulding (1966), Al Gore (2007). costs or externalities  positive or negative consequence of an activity by an actor and experienced by a third party without there being any kind of agreement between the parties (in the sense that the damaged person/beneficiary does not want the damage/benefit and the damage/benefit is not paid by those who produce it).

5 External

2.2 Environmental Economics

27

view, zero pollution would mean a zero level of economic activity: scientific literature reports on environmental economics refer that eliminating pollution is achieved only by stopping the production of the goods in question. We have to establish the optimal level of externalities so that the environment preserves its carrying capacity. In other words, if the level of waste R is lower than the level of assimilation A, the natural environment can carry waste, or pollution, and degrade and convert it into nondamaging or even useful products. This process will produce externalities (optimal level) (Pearce and Turner 1990). If R > A, the level of externalities is not optimal and regulations will be required to eliminate the surplus external costs. Emissions taxes are an example of tools to produce an efficient pollution level. The role played by this kind of tax is to incentivise an efficient allocation of resources rather than act as a tool/principle of justice, i.e., the polluter pays.

2.3 Tools and Principles of the Environmental Policy The Single European Act (1987) inserted an article (art. 36) about the environment into the Treaty of Rome; this gave the policy a formal juridical base and also established three main objectives: to preserve, protect and improve the quality of the environment; contribute towards protecting human health; and ensure a prudent and rational utilisation of natural resources. In 1992 the concept of sustainable development was inserted into EU legislation as part of article 130r, paragraph 2 of the Maastricht Treaty: “environmental protection requirements must be integrated into the definition and implementation of other Community policies”. This broadened the EU’s competence on the environment and in 1997 became one of the key objectives of the European Union in the Amsterdam Treaty. Even today, implementation of the environmental policy remains the most important future challenge for the EU which needs to focus primarily on making it much clearer by establishing more immediate operational mechanisms. Environmental policies act in synergy with the sustainable development strategies listed in the previous chapter. These strategies are generally divided into: • binding regulations and limits to be respected (prescriptive strategies). • incentives and competitive mechanisms (voluntary strategies). Compared to the seventies, the current tendency is to replace post-action tools (Polluter Pay Principle) with preventive methods (Prevention Cost and Willingness to pay) as well as the adoption of voluntary tools instead of Command and Control regulations.

28

2 The Environment and Economics

2.4 The Circular Approach to Design The relevance of the First Law of Thermodynamics was underscored in the paper entitled The Economics of the Coming Spaceship Earth written in 1966 by the economist Kenneth Boulding in which he represented the earth as a ‘SPACESHIP’. Boulding’s paper pointed to the need to consider earth as a closed economic system, one in which the economy and the environment are not characterised by linear linkages, but by a circular relationship. “The closed economy of the future might similarly be called the ‘spaceman’ economy, in which the earth has become a single spaceship, without unlimited reservoirs of anything, either for extraction or for pollution, and in which, therefore, man must find his place in a cyclical ecological system” (Boulding 1966). Just think of a spaceship going on a long journey; it will have only one external energy source: solar energy. It will have a stock of resources depending on whatever was put aboard before take-off. But as that stock is reduced, so the expected lives of the spacemen are reduced unless, of course, they can find ways to recycle water and materials and generate their own food. The spaceship is, of course, Earth. A similar vision was illustrated by Richard Buckminster Fuller in his “Operating Manual for Spaceship Earth” published in 1969. “Spaceship Earth was so extraordinarily well invented and designed that to our knowledge humans have been on board it for two million years not even knowing that they were on board a ship. And our spaceship is so superbly designed as to be able to keep life regenerating on board despite the phenomenon, entropy, by which all local physical systems lose energy. So we have to obtain our biological life-regenerating energy from another spaceship the sun” (Buckminster Fuller 1969). During the late seventies, John T. Lyle, a professor of landscape architecture, challenged graduate students to imagine a community in which daily activities were based on the value of living within the limits of available renewable resources without environmental degradation. During the next ten years, students and the faculty researched the possibilities of creating a community that used on-site resources operated with renewable energy, and worked with biologically-based processes. He founded the Cal Poly Pomona “Lyle Centre for Regenerative Studies”.6 Regenerative studies emphasise the development of community support systems that can be restored, renewed, revitalised, or regenerated through the integration of natural processes, community actions, and human behaviour. Some years later, the German chemist and visionary Michael Braungart and the American architect Bill McDonough went on to develop the Cradle to Cradle™ concept and certification process. This design philosophy considers all material involved in industrial and commercial processes to be nutrients, of which there are two main categories: technical and biological. Cradle to Cradle (C2C) design recognises the safe and productive processes of nature’s ‘biological metabolism’ as a model for developing a ‘technical metabolism’ flow of industrial materials. Product components can be designed for continuous recovery and reuse as biological and technical 6 http://env.cpp.edu/rs/rs

(access on 31/10/2018).

2.4 The Circular Approach to Design

29

nutrients within these metabolisms. The C2C design is based on the following three principles (McDonough and Braungart 2002): 1. Eliminate the concept of waste: “waste equals food.” Design products and materials with life cycles that are safe for human health and environment can be reused perpetually through biological and technical metabolisms. Create and participate in systems to collect and recover the value of these materials following their use. 2. Power with renewable energy: “use current solar income” to maximise the use of renewable energy. 3. Respect human and natural systems: “Celebrate diversity”. Manage water use to maximise quality, promote healthy ecosystems, and respect local impacts. Guide operations and stakeholder relationships using social responsibility. As the regenerative design approach, C2C is a biomimetic approach to the design of products and systems that models human industry on nature’s processes and considers materials as nutrients circulating in healthy, safe metabolisms. Designers and researchers active in different disciplinary fields explored the circular approach to design on both a global scale and an urban and architectural scale.7 This book focuses on this issue by developing a scalable and flexible design model that takes circularity into account in the life cycle.

References Boulding K (1966) The economics of the coming spaceship earth. In: Jarrett H (ed) Environmental quality in a growing economy. Johns Hopkins University Press, Baltimore, pp 3–14 Buckminster Fuller R (1969) Operating manual for spaceship earth, Southern Illinois University Press Duvigneaud P (1974) La synthèse écologique: populations, communautés, écosystèmes, biosphère, noosphere. Doin, Paris Falkie TV, McKelvey VE (1976) Principles of the Mineral Resource Classification System of the US Bureau of Mines and US Geological Survey, Bulletin 1450-A, US Government Printing Office, Washington Friedman Y (2016) Città immaginarie—villes imaginaires—imaginary cities. Quodlibet, Macerata Ghisellini P, Cialani C, Ulgiati S (2016) A review on circular economy: the expected transition to a balanced interplay of environmental and economic systems. J Clean Prod 114:11–32 Gore A (2007) An inconvenient truth, (dvd) Paramount Home Entertainment Guallart V (2010) Self-sufficient city: envisioning the habitat of the future. Actar, Barcelona McDonough W, Braungart M (2002) Cradle to cradle: remaking the way we make things, North Point Press 7 In

this field it is possible to cite several significant theories in urban and architectural experimentation: the concept of urban metabolism expressed, for example, in: the La synthèse écologique by Duvigneaud (1974); the Ville Spatiale (1958–2006) by Yona Friedman (Friedman 2016); the relationships between the environment and architecture characterised by the transformability and adaptability of one and the other; the concept of the Self-sufficient City defined by Guallart (2010) in which the city is an “urban habitat”, a multi-scalar system integrating environmental issue and urban planning. People, processes, information and flows of matter and energy are involved in the temporal sequence of the urban metabolism.

30

2 The Environment and Economics

Meadows D, Meadows DL, Randers J, Beherens W (1972a) I limiti dello sviluppo. Mondatori, Milano Meadows DH, Meadows DL, Randers J, Behrens WW III (1972b) The limits to growth. Potomac Associates—Universe Books, New York Musu I (2000) Introduzione all’economia ambientale, il Mulino, Bologna Panella G (2002) Economia e politiche ambientale. Carocci, Roma Pearce DW, Turner RK (1990) Economics of natural resources and the environment. The Johns Hopkins University Press, Baltimore

Chapter 3

Life Cycle Methodologies

Abstract This chapter introduces two methodologies based on the life cycle concept: Life Cycle Assessment (LCA) and Life Cycle Costing (LCC). LCA and LCC are briefly described in order to provide the reader with an overview of the procedures and a complete bibliographic framework. The first two sub-chapters focus on the origins, standards, studies, references and methods used to calculate the life cycle approach to buildings. Part three combines LCA and LCC in order to define a common framework that can be used to develop the Life Cycle Model described in Chap. 4. It includes an outline of ongoing projects based on the combination of the two methods. The decision to present these evaluation techniques focusing on the building analysis illustrated early on in the book was inspired by the need to establish the cultural and scientific background for the experimental part of the monograph in Part II.

3.1 Life Cycle Assessment (LCA) 3.1.1 Origins, Definitions, and References The first time the “from cradle to grave” slogan of the Life Cycle Thinking approach1 was applied to environmental assessments occurred in the sixties in the United States; it was chiefly sponsored by big companies and the Environmental Protection Agency (EPA). The LCA analysis was originally known as the Resource and Environmental Profile Analysis (REPA); it was developed to perform energy-environmental assessments 1 Life Cycle Thinking defines the principles used to ensure the continuous improvement of environmental performance at every stage of the life cycle of a system: from design (eco-design), to production, business management (EMAS and ISO 14000), disposal, and end of life. The strategies used to assess sustainability are based on the Life Cycle Management principle (LCM) according to which the life cycle and economic, environmental, and societal considerations are integrated into the decision-making processes regarding product development. (UNEP/SETAC Life Cycle Initiative, LCM Definition Study in Saur et al. 2003)

© Springer Nature Switzerland AG 2019 F. Thiébat, Life Cycle Design, PoliTO Springer Series, https://doi.org/10.1007/978-3-030-11497-8_3

31

32

3 Life Cycle Methodologies

of the life cycle of different types of industrial food packaging. The objective was to compare the environmental impact of different solutions and materials with the same functions. At the time this approach was a complete novelty: the industrial system had always focused on individual components, but from this moment on it started to concentrate on the whole production chain. In Europe, the English researchers Boustead and Hancock published the Handbook of Industrial Energy Analysis (EA) in 1979.2 Their study was the first to lend operational support to the analytical method; many experts still consider it a crucial text for the current LCA (Baldo et al. 2005). During that same period, the oil crisis, the dawn of the environmentalist movement, environmental economics, and the advent of the sustainable development concept created the humus used by scientists all over the world to develop multiple theories and studies on environmental impact issues: pollution and resource consumption. The term Life Cycle Assessment (LCA) was coined during the congress organised by the Society of Environmental Toxicology and Chemistry (SETAC)3 held in 1990 in Smuggler Notch (Vermont, USA). The congress focused on uniforming the studies performed up to that point. In the early nineties, numerous standardisation initiatives began to be implemented in parallel with the increased use of manual calculation tools and databases to encourage the practical use of the LCA. The first standard series is the ISO 14040/44 on Environmental Management—Life Cycle Assessment published in 1998, revised in 2006 and confirmed in 2016. The standard had been drafted by the ISO/TC 207 team (Fig. 3.1). During that period, new environmental policies based on Life Cycle Thinking were also developed in Europe.4 The methodological approach in all the proposed environmental policies is based on LCA studies highlighting the problematic environmental issues affecting the system. Even if the ISO guidelines on Life Cycle Assessment defined a general framework, they did not provide a technically-detailed standardisation (EC, JRC, IES 2010). The UNEP-SETAC Life Cycle initiative worked toward consensus and recommended best practices; over time it has been complemented by the work of many other organisations, such as the Environmental Protection Agency of the United States (US EPA) and the European Commission (i.e. the Joint Research Centre—JRC). Since 2010 the European Commission has published a set of technical guidance books for the application of ISO 14040/44 standards: the International Reference Life Cycle Data System (ILCD). The handbooks provide governments and businesses with a basis for ensuring quality and consistency of life cycle data, methods and assessments (Wolf et al. 2012; EC, JRC 2010). 2 Boustead

and Hancock (1979). is a professional company for environmental sciences and engineering and correlated disciplines interested in the quality of the environment. It has offices in Florida (USA) and Belgium. 4 For example, several tools are used to direct the market towards eco-responsible choices in order to encourage the demand for low environmental impact products and services, including: Integrated Product Policies (IPP), Green Public Procurement (GPP), Ecological labels (Ecolabel, EPD), and integrated waste management. 3 SETAC

3.1 Life Cycle Assessment (LCA)

33

Fig. 3.1 ISO-TC207—environmental management (SC 3, SC 5 and SC 7). The table shows some of the standards used to develop the LCDM described in this book Fig. 3.2 Diagram of the structure of the LCA based on ISO 14040

The UNI EN ISO 14040 standard defines the LCA a “compilation and evaluation of the inputs, outputs and potential environmental impacts of a product system throughout its life cycle”. The inputs of the system are the parameters that intervene in the debate about the problems regarding energy resource saving, while the outputs involve pollution and waste problems. The LCA defines the system that generates the product and how it functions. Unlike the structure theorised by SETAC in 1990, the one proposed by the ISO can be summarised in four main stages (Fig. 3.2).

34

3 Life Cycle Methodologies

1. Goal and Scope Definition The objectives of the analysis are identified and described during this initial stage together with the field of application, the functional unit, and the boundaries of the LCA study. The latter establishes the general approach of a LCA study; it describes the system to be studied and identifies the categories of the required data as well as its assumptions and limits. 2. Life Cycle Inventory Analysis—LCI This stage includes data collection and calculation procedures to quantify the inputs and outputs of a system. The sequential processes of a production system are established during this stage; it involves identification of the required quantities of energy and raw materials in order to reproduce a theoretical model that can represent the functioning of the real system (flow chart). This is undoubtedly the most important and complex stage of a LCA study. For this reason dedicated software and databases are generally used during this stage. 3. Life Cycle Impact Assessment—LCIA The results of LCI are processed during this stage in order to evaluate the level of potential environmental impacts, highlight the extent of the environmental alterations generated by emissions or waste, and assess the resource consumption caused by production. Processing involves classifying, characterising, and normalising the data according to the contribution it can make to the formation of potential environmental effects. Impact analysis entails turning the objective inventory analysis into an environmental evaluation based on investigative elements updated over a period of time and subject to uncertain variations. 4. Life Cycle Interpretation This is the final stage of the life cycle assessment during which the results of the inventory and the impact assessment analysis are consistently combined with the pre-established goal and scope of the study. The objective of the interpretation stage is to arrive at conclusions and recommendations to reduce the environmental impact of the processes or activities in question, evaluating them iteratively with the same methodology. Several international and European work groups developed frameworks and methodologies for the construction industry in line with ISO TC 207, for example ISO-TC59-SC17 and ISO-TC59-SC14 “Buildings and civil engineering works” and CEN/TC 350 “Sustainability of Construction Works”.5 The goal was to promote the harmonised application of tools and methods to make new or existing buildings environmentally, economically and socially sustainable, thereby satisfying the three pillars of sustainability.

5 Cfr.

Chap. 1.

3.1 Life Cycle Assessment (LCA)

35

3.1.2 Environmental Indicators The choice of impact indicator is crucial because it has to establish the effects that need to be considered in line with the objectives of the analysis. Each environmental effect corresponds to a calculation method and an impact indicator. The Life Cycle Impact Assessment phase (LCIA) of Standard 14044 requires mandatory: • Selection of impact categories • Definition of the elementary flows from the life cycle inventory, then assigned to impact categories (“classification”) • Calculation of the impact of each emission or resource consumption, expressed as an impact score in a unit common to all contributions (characterisation factors) within the impact category (“characterisation”).6 “Normalisation” and “weighting” are optional; the former associates impact scores with a common reference to facilitate comparison of environmental indicators; the latter determines the importance of the different environmental impact categories. According to the study by SECTAC and indications by the CEN TC/350, the environmental variable can be summarised in one or more indicators of the environmental effect, in other words it can be represented by just one impact indicator if the latter is crucial as regards the total environmental weight of the analysed product. CEN TC/350 does not include the normalisation and weighting process which are optional in the ISO 14040/44. To determine the sustainability indicators for buildings one can refer to Standard ISO 21929-1 Sustainability in building construction—Sustainability indicators—Part 1: Framework for the development of indicators and a core set of indicators for buildings (TC59/SC17). Based on the general principles of sustainability, this standard provides a framework for the development and use of indicators to evaluate the economic, environmental, and societal impacts of buildings. The standard follows the general principles established by ISO 15392 and is congruent with the family of international standards defined by TC/207 (Fig. 2.1, ISO 14020, 14040, and 14067). An analysis of the environmental data of most building products7 shows that an environmental assessment for the building sector focuses more often on two impact indicators: the first is the Global Warming Potential (GWP) that evaluates the greenhouse gas effect, while the second is the consumption of primary energy resources identifying energy flows from non-renewable and renewable sources. Selection of these indicators is determined by several important factors common to both, including: 6 The

“midpoint indicators” or “impact indicators” are expressed through the characterisation process, while the “endpoint indicators” express the categories of damage and require a normalisation process. 7 Data regarding characteristics was taken from scientific literature (Giordano 2010; Blengini and Di Carlo 2010; Lavagna 2008), software processing programmes (Sima-Pro), databases (Boustead 4.0, Ecoinvent, I-lca, Buwal), and handbooks.

36

3 Life Cycle Methodologies

• they have a global effect; • they use internationally acknowledged characterisation factors such as polluting substances or non-renewable resources, thereby significantly limiting the subjective component often leading to uncertain results and making the results easier to understand for users unfamiliar with LCA; • they are included in the field of environmental economics (i.e., pollution and scarce resources); • they can be compared with the energy simulation assessment indicators for the operational stage of buildings (i.e., Energy Performance of Buildings Directive—EPBD) and with the reference standards used to design nearly Zero Energy Buildings (nZEB). The two aforementioned indicators, and several indicators that may be representative of certain special situations, are presented below.8

3.1.2.1

Global Warming Potential

While several gases in the atmosphere can be crossed by the sun’s short waves, they withhold the longwave infrared radiations reflected by the earth’s surface, thereby ensuring suitable temperatures here on earth. An increase in the concentration of the gases capable of absorbing and re-emitting the infrared radiation emitted by the earth (e.g., CO2 and water vapour) is called the “greenhouse effect”. The latter contributes to the global warming of the earth’s atmosphere and ensuing changes in the climate. Despite the fact there is no linear link between the greenhouse effect and climate change, many studies maintain that excessive emissions of these gases produced by human activities have altered the energy balance and caused an increase in temperature (Global Warming). Two agencies of United Nations (UN) were merged in 1988 to create the Intergovernmental Panel on Climate Change (IPCC): the World Meteorological Organization (WMO) and the United Nations Environment Programme (UNEP). The task of the IPCC was to study global warming.9 The Global Warming Potential (GWP) is the midpoint indicator of the impact category, while the characterisation factor is represented by the equivalent kilograms of carbon dioxide. Based on concentration and period of exposure, the GWP measures the potential input of a substance into the greenhouse effect, compared to the input caused by the same amount (weight) of carbon dioxide. The exposure period or “time-horizon” varies; generally speaking, a 100-year GWP is considered since the longer the period, the more uncertain the estimates. GWP values are established by the IPCC as indicated in Table 3.1. 8 For example, in Sect. 5.2—the “Chavonne warehouse” case study—other environmental indicators

were considered because they were deemed important in order to evaluate the effects of biomass combustion (Acidification Potential and Eutrophication Potential). 9 The last Special Report published by the IPCC on Global Warming is dated October 2018 (access 31/10/2018).

3.1 Life Cycle Assessment (LCA)

37

Table 3.1 The global warming potential of several greenhouse gases: conversion factor in kg of CO2 eq according to the IPCC (values in the table do not include climate-carbon feedbacks). The 2014 values (AR4) are the most recent. For a complete list of the gases, refer to IPCC publications (http://www.ipcc.ch—last access 24/10/2018) Gas name

Carbon dioxide

Chemical formula

CO2

GWP values for 100-year time horizon IPCC second assessment report SAR

IPCC fourth assessment report AR4

IPCC fifth assessment report AR5

1995

2007

2014

1

1

1

Methane

CH4

21

25

28

Nitrogen oxide

N2 O

320

298

265

Sulphur hexafluoride

SF6

23,900

22,800

23,500

CFC-11

CCl3 F

3800

4750

4660

HFC-134a

CH2 FCF3

1300

1430

1300

PFC-14

CF4

6500

7390

6630

Unlike other atmospheric pollution phenomena that cause an impact globally, e.g. acidification and eutrophication, due to substances such as SOx (sulphur oxide), or NOx (nitrogen oxide), or an impact regionally (ozone and PM10), or locally (CO and benzene), the GWP involves the whole planet and should therefore be considered a global phenomenon. The following are the most important greenhouse gases: Carbon dioxide (CO2 ): Carbon dioxide is considered the most important greenhouse gas and the secondlargest contributor to the greenhouse effect after water vapour. All activities linked to the consumption of fossil fuels cause the CO2 stored millions of years ago to be released into the atmosphere. Deforestation also contributes to an increase in emissions (roughly a third) since on the one hand it involves the release of CO2 stored in trees and, on the other, decreases the absorption that would have taken place through chlorophyll photosynthesis. Several criteria exist in literature to calculate carbon dioxide emissions from biogenic sources. In fact, carbon dioxide credits must be assigned when considering the use of vegetal material, i.e., the quantity of carbon dioxide that woody biomasses capture from the atmosphere during growth (e.g., biomass can be used as a building material or as fuel). The first criteria envisages prior assignment of the CO2 credit to woody biomass input into the LCA model, after which, during the end-of-life stage all carbon dioxide emissions, including from biogenic sources, are included in the calculation of the GWP indicator, with no exceptions. The second criteria does not envisage prior assignment of the CO2 credit, but the biogenic emissions of carbon dioxide are not included at end of life (they are considered neutral from the point of view of the greenhouse effect). The third criteria envisages post assignment of the

38

3 Life Cycle Methodologies

CO2 credit only to the biomasses which at the end of life of the building are reused after calculation of the biogenic emissions. When a rigid LCA model is created (i.e., from cradle to grave), the three aforementioned criteria produce the same result. In short, as often happens when the model is applied to real cases, the first criteria can lead to the assignment of CO2 credits that are not “returned” at the end of life, thereby altering the results of the analysis. Hence, the more conservative second and third criteria should be adopted because they assign CO2 credits only to the biomasses acting as a carbon dioxide “storage” that permanently captures the carbon dioxide (Blengini and Di Carlo 2010). Water vapour: Water vapour is the most abundant greenhouse gas in the atmosphere and the most efficient when it comes to trapping the heat emitted by the earth (its contribution to the greenhouse effect is estimated at 60%). Nevertheless, since human activities seem to produce very small changes in its concentration, water vapour is not considered a “real” greenhouse gas in studies on man’s activities, but only in studies on climate change (Baldo et al. 2005). Methane CH 4 : Even if the concentration of CH4 is lower than that of CO2 , its radiative power is twenty times superior (Table 2.1) and it is responsible for 20% of the increase in the greenhouse effect. It is produced by the bacteria responsible for organic decomposition, by waste dumps, and animal species. It is also released during the production and transportation of carbon and natural gas. It stays in the atmosphere for only eleven to twelve years and is captured in the natural water formation process. The greenhouse effect is a natural phenomenon needed to regulate the warming of the earth. In fact without it the average temperature on earth would drop by 33 °C, i.e., it would be—18 °C against the current 15 °C (Baldo et al. 2005). The industrial revolution and subsequent massive use of fossil fuels has caused a substantial increase in the concentrations of anthropic greenhouse gases such as carbon dioxide, methane, nitrogen oxide, and sulphur hexafluoride. The earth’s capacity to trap heat has risen enormously, causing changes in both the local and global climate. According to studies by the IPCC, the continuous rise in emissions and concentration of greenhouse gases could raise the earth’s temperature by 1.4–5.8°C in the next century, causing climate changes that are difficult to forecast. These changes could be linear, caused by current trends, or sudden and non-linear; the latter are called abrupt changes (Baldo et al. 2005, p. 171). According to the last IPCC Special Report published in 2018, if we want to limit global warming by 1.5 °C—as envisaged by the last Conference of the Parties (COP21, Paris Agreement 2015–16)—we need to implement rapid, farsighted transitions in many sectors such as land, energy, industry, building, transportation and urban planning. The net, global CO2 emissions produced by human activities should drop by roughly 45% before 2030, compared to the 2010 levels, and reach zero around 2050. Every remaining emission should be balanced by the removal of CO2 from the atmosphere (e.g., using Carbon Capture and Storage technologies).

3.1 Life Cycle Assessment (LCA)

39

The Carbon Footprint category indicator linked to the GWP can be defined as “a measure of the total amount of emissions of carbon dioxide (CO2 ) and methane (CH4 ) [and other Greenhouse gases (GHGs) in the context of LCA] of a defined population, system or activity, considering all relevant sources, sinks and storage within the spatial and temporal boundary of the population, system or activity of interest. Calculated as carbon dioxide equivalent using the relevant 100-year global warming potential (GWP100)” (Wright et al. 2011). The Carbon Footprint standard was published in 2013 and revised in 2018: ISO 14067 Greenhouse gases—Carbon footprint of products—Requirements and guidelines for quantification. This standard, consistent with International Standards on life cycle assessment (LCA) (ISO 14040 and ISO 14044), specifies principles, requirements and guidelines for the quantification and reporting of the Product Carbon Footprint (PCF). Embodied Carbon/Operational Carbon Embodied Carbon (EC) is an impact indicator based on the potential quantification of the concentration of carbon dioxide released during extraction of raw materials, product transformation, and transportation. It is applied to building materials by either considering the system boundary “from cradle to gate”, i.e., up to the factory gates, or by extending the boundary “from cradle to site” (Hammond and Jones 2008). The difference between EC and GWP is the fact that the former refers only to preconstruction emissions, while the latter refers to the “from cradle to grave” life cycle, including the operational and disassembly stages of the building and materials. It is measured in kg of CO2 eq and can be integrated with data from the energy simulation of the use stage. Embodied Carbon (EC) refers to the proportion of CO2 embodied in building materials. It can be added not only to Operational Carbon (OC) which is the proportion of CO2 associated with the energy performance of the building, but also to Recurring Embodied Carbon (R-EC). The latter refers to the proportion of CO2 in building materials used during maintenance/replacement.

3.1.2.2

Resource Consumption (Energy and Materials)

The energy flows calculated in the LCA is the direct energy consumed during the whole process, while the material flows are the ones used to create it. In particular, the resources are divided into renewables (air, water, and sunlight) and non-renewables (mineral resources and the territory). In addition, their current consumption is compared to the consumption of the total ascertained resources. The Gross Energy Requirement (GER), expressed in MJ or kWh, is the total primary energy required to produce a product.

40

3 Life Cycle Methodologies

The GER is the overall energy that has to be extracted from nature. It represents the sum total of direct energy (energy consumed directly by process operations), indirect energy (energy used by the energy industry to produce and transport direct energy), feedstock energy (energy contained in materials—for example the oil in plastics, biomass, etc.) and Capital Energy (energy to build machinery and infrastructures). One can choose whether or not to include the feedstock energy stored in materials in the sum total of energy. Like the CO2 absorbed by biomass, feedstock energy can in fact be considered as stored in materials until it is used (e.g., in combustion). To avoid double calculations, it should preferably be calculated only if the material is a fuel, or if the system boundary of the LCA analysis envisages energy recovery of the materials at end of life. Based on the calculation method used to measure the impacts of the LCIA, the energy indicator can be established using other indicators similar to the GER, for example Primary Energy Impact (PEI), Embodied Energy (EE), Operational Energy (OE), and Cumulative Energy Demand (CED). Although in practice designers tend to focus more on calculating the energy balance of a building based on its operational energy consumption (i.e., Directive 2010/31/EU), there is growing interest in the need to consider the whole life cycle, with LCA assessments, by integrating energy assessments during the use stage according to the nZEB approach (Thormark 2002; Bonanomi et al. 2012; Deng et al. 2014, Giordano et al. 2015; Monticelli and Thiébat 2016). The project entitled Towards Net Zero Energy Solar Buildings (NZEBs)10 promoted by the International Energy Agency (IEA) considers embodied Energy (EE) as one of the possible indicators that can be used to evaluate the energy impact of buildings (Voss and Musall 2012; Karsten and Riley 2009; IEA 2014). Embodied Energy refers to the proportion of energy incorporated in building materials that can be integrated with data from the energy simulation of the use stage represented by the Operational Energy (OE) indicator, i.e., the proportion of energy associated with the energy performance of the building.11 In addition, it can be added to the proportion of the building materials used during building maintenance/replacement: Recurring Embodied Energy (R-EE).

3.1.2.3

Other Environmental Impact Indicators

Other indicators can be included if they are relevant for specific goals. In particular, reference is made to the aforementioned standards framework and the methods used to characterise the LCIA impacts.12 Some examples of environmental impact indicators are provided below: 10 IEA

SHC TASK 40—IEA ECBCS ANNEX 52. the existing certification methods, for example MINERGIE-ECO (the Swiss protocol for sustainable building) envisages the inclusion of the embodied energy of building materials and end-of-life processes and disposal in the assessment of the energy consumption of the building. 12 The list is incomplete. Regarding the choice of environmental impact indicators, for buildings refer for example to: the guidelines of the Environmental Declaration Product for product assessment 11 Among

3.1 Life Cycle Assessment (LCA)

41

Abiotic depletion potential (ADP-elements) for non-fossil resources equiv kg Sb This category is linked to the protection of the wellbeing of humans and, in general, ecosystems. It relates to the extraction of mineral and fossil fuels required to satisfy the energy needs of the analysed system. The Abiotic Depletion Potential (ADO) is measured by an index obtained from the ratio between kilograms of equivalent Antimony Sb and kilograms of extracted fuels. It is a global indicator (i.e., CML method, Guinée et al. 2002). Acidification potential of soil and water, AP equiv kg SO2 This is a process that converts the polluting substances in the atmosphere into sulphuric and nitric acids, including ammonium, sulphur oxide SOx , and nitrogen oxide NOx released by the combustion of oil, biomass, and agriculture. The acidifying substances released in acid rain have a huge impact on the soil, surface water, water tables, living organisms and their ecosystems, and outdoor infrastructures (buildings, roads, etc.) The Acidification Potential (AP) for emissions into the air can take into account the origin and destination of the emitted acidifying substances. AP is expressed as equiv kg SO2 /kg emission (i.e., CML method, Guinée et al. 2002). Ozone depletion potential of the stratospheric ozone layer, OPD equiv kg CFC 11 The ozone layer in the stratosphere fifteen to sixty kilometres above the earth is made up of oxygen. It protects the earth from shortwave ultraviolet radiation (UV-B). In the late eighties a group of scientists discovered that two holes were forming in the ozone layer in the Artic and Antarctic due to its interaction with the chlorine oxides in chlorofluorocarbon gases (CFC), hydrochlorofluorocarbons (HCFC) and hydrofluorocarbons (HFC). Increasing quantities of UV-B radiation produces harmful effects on the health of humans, animals, soil and water ecosystems, biochemical cycles, and materials. Characterisation defines the potential reduction of the ozone layer of several gases as a ratio between the emitted kg of equivalent CFC-11 and the total emissions produced (i.e., CML method, Guinée et al. 2002). Eutrophication potential, EP equiv kg (PO4 )3 Eutrophication (also called Nutrification) includes all the impacts caused by the excessive presence of macronutrients in the environment (nitrogen and phospherous) due to their release into the air, water, and soil. They are present primarily in the fertilisers used in agriculture and industrial and urban dumps. (e.g., Product Category Rules—PCR—for the assessment of the environmental performance of UN CPC 531 Buildings); the characterisation methods of LCIA impacts to calculate the indicators of single impacts, midpoint indicators, or aggregates, endpoint indicators (e.g., CML, ReCiPe, EcoIndicator, etc.); international standards ISO 21929 establishing the building indicators; other specific design references.

42

3 Life Cycle Methodologies

The nitrification potential (NP) is based on a stoichiometric procedure defined by Heijungs in 1992 and expressed in equivalent kg PO4 for each kg emitted (i.e., CML method, Guinée et al. 2002). Photochemical ozone creation potential of tropospheric ozone, POCP equiv kg Ethene C2 H4 The impact category, often indicated as summer smog, is associated with the formation of reactive substances (especially ozone in the lower strata of the atmosphere) that are harmful for humans and ecosystems in general and can damage agriculture. The Photochemical Ozone Creation Potential (POCP) is expressed in equivalent kg of ethylene for each kg emission (i.e., CML method, Guinée et al. 2002). Human toxicity This measures the effects of toxic substances on human health. The characterisation factor, human toxicity potential (HTP), was calculated taking into account human life expectancy and exposition to and effects of toxic substances. For each toxic substance the HTP is expressed by the ratio between the quantities of equivalent 1,4dichlorobenzene and the kg of emissions (i.e., CML method, Guinée et al. 2002). Fresh-water aquatic, Marine and Terrestrial eco-toxicity Impact on several ecosystems due to the emission of toxic substances in the air, water, and soil. The eco-toxicity potential (FAETP) can take into account human life expectancy and the exposition to and effects of toxic substances. The characterisation factor is expressed by the ratio between the quantities of equivalent 1,4dichlorobenzene and the kg of emissions (i.e., CML method, Guinée et al. 2002).

3.2 Life Cycle Costing 3.2.1 Origins, Definitions and References Life Cycle Costing (LCC) is a long-standing technique. In the United States life cycle costing was included in maintenance and management costs back in 1933 when the General Accounting Office (GAO) was purchasing agricultural tractors. It’s interesting to note how even eighty years ago people understood that in procurement contracts there was far too much emphasis on purchase costs compared to management and maintenance costs. The latter, in fact, can easily be much higher than the initial investment. In the seventies the US government made Life Cycle Costing obligatory by law for weapons contracts and building programmes by public institutions in many states in the US (Hunkeler et al. 2008). Since the early seventies the US Department of Defence drafted numerous directives to calculate the Life Cycle Costing of extremely expensive military equipment such as airplanes and tanks; these directives included the life cycle cost of a product or system in the R&D stage. Investments were not to be based only on the initial purchase cost, but also include operational and maintenance costs and, to a lesser

3.2 Life Cycle Costing

43

degree, disposal costs. The need to allocate an optimal budget for the life cycle of a system, product, or technical element vis-à-vis its economic performance is another reason corroborating the importance of Life Cycle Costing (Hunkeler et al. 2008). Initially the LCC was only an economic tool, aimed at analysing past, present and future costs in order to choose the most cost-effective option. As Glunch and Baumann (2004) rightly mention, traditional LCC does not become an environmental tool just because it contains the words “life cycle”. Based on this tradition, the current LCC is applied primarily to decisions regarding the purchase of equipment or long-lasting products with high investment costs per unit in the following fields: • Expensive real estate investments (especially, public or commercial buildings) • Use and production of energy • Transport vehicles with a high investment cost (especially for the aerospace industry) • Important military equipments and arms. Several regulations have been published in the US requiring Life Cycle Costing to be calculated during the purchase of public buildings. Nevertheless, LCC has usually been limited to specific applications for a product or sector. The book published in 1981 by Sherif and Kolarik provides a complete panorama of all the applications of the cost assessment models used so far, as well as reference literature.13 They note that “Life Cycle Costing was developed more as a result of specific applications rather than hypothetical models”. This conclusion still holds true. In 2005, Gerald Rebitzer, an engineer specialised in the implementation of product sustainability, maintained that up till then no model or widely usable methodological framework had been developed, although many attempts had been made in this direction. The concept most similar to the general LCC method was initially developed by Blanchard14 in 1978 and then fine-tuned by Blanchard and Fabrycky in 1998. Several guidelines/standards were developed in this direction, including: EU Task Group 4 Final Report (published by the European Union in 2003), IEC 60300-3-3 (International Electrotechnical Commission 2004), ASTM E917-5 (US Standards for Guidance 1999) and finally AS/NZS 4536 (Australia and New Zealand Standards for Guidance 1999). In the building sector, standards EN 16627 (Sustainability Assessment of Buildings—Economic Performance of Buildings. Published in 2015) and ISO 15686-5 (Building and constructed assets—Service life planning—Life Cycle Costing. Published in 2008 and updated in 2017) are the most recent tools for the application of 13 Sherif, Kolarik “Life Cycle Costing: Concept and Practise” in Omega. The International Journal of Management Science. Reference from the book by Hunkeler D., Lichtenvort K., Rebitzer G., Environmental Life Cycle Costing, SETAC Books, CRC Press, New York, 2008. 14 Blanchard’s articles were published in the book by Hunkeler D., Lichtenvort K., Rebitzer G., Environmental Life Cycle Costing, SETAC Books, CRC Press, New York, 2008. However, the authors maintain that Blanchard and Fabrycky, and the standards, do not elaborate a methodology, but provide guidance about how to calculate and compare costs. They present LCC more in terms of Life Cycle Thinking and stress the importance of the global vision of the system.

44

3 Life Cycle Methodologies

LCC in the building industry. The engineering of industrial systems gave birth to the regulatory system focusing on the assessment and comparison of technological alternatives. The reference framework of the life cycle of a product or a system is normally divided into stages: R&D, production and construction, use and maintenance, end of life and disposal. This framework is behind the Life Cycle Thinking approach. For many years the Society of Environmental Toxicology and Chemistry (SETAC) has worked on global cost as part of Life Cycle Thinking issues. It maintains that the difficulties inherent in the application of a standardised LCC methodology for common applications are due to the fact that in many cases it does not reflect the cost assessment system used by individual companies or organisations. Since a LCC methodology must be reflected in these assessment systems, any attempt to provide a universally valid application means that the specific cost values of individual companies have to be turned into more generic and therefore less efficient values (Hunkeler et al. 2008). Taking all this into consideration, we can say that SETAC, along with voluntary international standardisation organisations such as ISO and CEN, producer associations, universities, and other research agencies, are all trying not only to find optimal solutions to overcome the problems regarding economics and management, but, more in general, to deal in an integrated manner with the problem of sustainable development and its three “pillars”: environmental, economic, and societal. In the construction industry Life Cycle Costing caught the attention of the public sector in the mid-seventies. In Europe there is a growing demand for use stage costs to be included in public projects and building procurement contracts (Glisoni et al. 2010; Estevan et al. 2018). For example, in Italy the current public contracts code Nuovo Codice Appalti (Decree Law 18 April 2016, n. 50) implementing Directives 2014/23/EU, 2014/24/EU and 2014/25/EU, establishes that Life Cycle Costing be included in contract awarding criteria (art. 95, II comma and art. 96).15 Apart from the programming and planning of sustainable building interventions, the Global Cost concept is an integral part of the regulatory framework as regards measurement of the energy performance of buildings using, for example, the Cost Optimal concept.16 In particular, the latter sets the minimum energy performance level required for buildings, taking into account their Energy Class (Fregonara 2015, 2017). 15 The life cycle approach had already been introduced into European policies in Decision n. 1600/2002/EC stating that “this requires promoting a green public procurement policy, allowing environmental characteristics to be taken into account and the possible integration of the environmental life cycle” (art. 3.6) and in the Communication from the Commission to the Council and European Parliament “Integrated Product Policy. Building on environmental life-cycle thinking” COM(2003)302. 16 The relationship between sustainability and energy efficiency is assumed by Directive 2010/31/EU with guidelines n. 244/2012 which establish a comparative methodology framework for calculating cost-optimal levels of minimum energy performance requirements for buildings and building elements. Furthermore, standard EN 15459-1—Energy performance of buildings—Economic evaluation procedure for energy systems in buildings—Part 1: Calculation procedures, Module M1-14 provides a calculation method for the economic issues regarding heating systems and other systems.

3.2 Life Cycle Costing

45

The LCC in the service life cycle was invented well before the concepts of sustainable development and Life Cycle Thinking. It was used to perform technicaleconomic analyses in the industrial sector and was later applied in the construction industry to compare alternative design strategies during the following stages: • design stages of new buildings (life cycle cost planning) as regards initial capital costs and envisaged future costs in the service life of the building and its elements • transformation and recovery stages to select the elements and materials to be used in the building according to their economic affordability in the life cycle • maintenance stages to optimise the threshold of affordability of replacing an element or a component already in use. This kind of analytical model requires available data regarding the maintainability and durability of the technical elements and operational costs over a period of time. Maintainability means “the aptitude of an entity in certain conditions of use to be maintained or restored to a state in which it can perform the required function, when maintenance is performed in given conditions, with prescribed procedures and means” (UNI 9910). Durability and service life mean “the ability of a building, an assembled system, a component, a product, a construction to maintain a state of efficiency for at least a defined period of time” (CIB W80—RILEM 7117 —Prediction of service life). However, while in industrial production the obsolescence of products must be planned, managed and programmed, in the building sector, maintenance has often been considered in the previous design phase as an accessory activity of little value (Pollo 2015). Today, anticipating maintenance activity, related costs and environmental impacts is mandatory for architects focusing on a life cycle design perspective. Mixed cost estimate procedures that combine analytical methods (i.e., Bill of Quantities) with synthetic methods (i.e., parametric estimate) are well suited to the LCC application during different design stages, including the initial stages and the disassembly of a building into its component parts for new buildings or restoration or maintenance projects. Mixed procedures can be performed (Fregonara 2015): • through disassembly of building elements: i.e., either as the sum of the cost values of certain building elements of the building, or by parametric values due to similarity with comparable buildings, or by an analysis dictated by the cost of every building element; • according to typologically homogeneous samples: adaptation of the quantities of known cost values of a similar building type; The disassembly of building elements can be based on common use patterns18 (i.e. load-bearing structures, closure/envelope, internal walls, external walls and connecting systems, technological systems, furnishings, equipment, and external layouts). 17 CIB (Conseil Internazionale du Batiment) and RILEM (Réunion Internationale des Laboratoires et Experts des Matériaux, systèmes de construction et ouvrages). 18 For example, in Italy it is possible to use the class division of technological units defined by the UNI 8290 standard.

46

3 Life Cycle Methodologies

Up until 2008, the reference framework regarding the application of LCC to the real estate sector was still very fragmented, broken down sector by sector, and based chiefly on reference studies. Later on the publications by SETAC, ISO and CEN provided a relatively transparent or at least sufficiently coherent methodological framework, in particular as regards the three tools on which this monograph is based: 1. in June 2008 the International Standards Organisation published ISO 15686-5 “Buildings and Constructed Assets—Service Life Planning—part 5—Life Cycle Costing” regulating application of the method as regards durability of buildings; 2. in 2008, after five years of team work, the SETAC Working Group on LCC proposed the Environmental LCC, a method linking LCA and LCC; the method is not specific to the building sector, but enables the LCC application to be used in several sectors; it was summarised in a Code of Practice published in 2011; 3. in 2012, the CEN TC 350 published EN 15643-4 Sustainability of construction works—Assessment of buildings—Part 4: Framework for the assessment of economic performance and in 2015, EN 16627—Sustainability of construction works—Assessment of economic performance of buildings—Calculation methods as the result of WG4 focusing on economics in the assessment of the sustainability performance of buildings. An analysis of the three tools made it possible to identify a general framework, a common methodology, and a glossary applicable to economic assessments in the field of construction, based on the one adopted for LCA by ISO 14040/44 standards.

3.2.1.1

Standards: ISO/TC59/SC14 and CEN/TC350

In June 2008 the International Standardisation Organisation (ISO) published ISO 15686-5 which was revised in 2017. The standard was developed in collaboration with the International Electrotechnical Commission (IEC) and drafted by the technical committee ISO/TC 59 (Building Construction) SC 14 (Design life). The objective of the ISO 15686 standards series Building and constructed assets—Service life planning is to define the service life planning concept19 for new or existing buildings. The series is based on the documents published by the Conseil Internazionale du Batiment (CIB), the Réunion Internationale des Laboratoires et Experts des Matériaux, systèmes de construction et ouvrages (RILEM), and also case studies and applications performed in several countries (UK, USA, Canada, Japan, etc.).

19 The term indicates the planning of the envisaged life of the building and its components during the design stage. For example, the goal of the service life planning could be to reduce ownership costs and facilitate management and maintenance processes. Several terms are also linked to the “useful life” concept: predicted service life—predicted life according to technical production specifications design life—prediction by the designer service life—prediction based on statistical data of already installed elements.

3.2 Life Cycle Costing

47

For the purposes of this study, Part 5 is the most representative. Its objectives can be summarised as follows: • Establish comprehensible terminology and a common methodology for Life Cycle Costing • Allow practical use of the method in order to promote it within the construction industry • Provide instructions to perform a LCC assessment suited to the project type • Improve the decision-making and assessment processes during the most important stages of any project • Define the differences between Life Cycle Costing (LCC) and Whole Life Costing (WLC) • Provide a generic list of costs, for LCC/WLC, that can be personalised and compatible with national and international standards and particular company practices. Bearing this in mind, standard 15868-5 defines LCC as a useful method to estimate and quantify the economic performance of buildings and building components over a period of time. It is an efficient analytical tool to verify correspondence between design choices and clients’ requirements. Unlike Life Cycle Costing, which includes costs such as construction, operation, use, and end of life, Whole Life Costing also considers non-construction costs, revenue, and externalities. Externalities can be either positive or negative depending on the object: benefit or damage. They embody the benefits allocated or damages caused to third parties by an entity in the course of its activity without there being an agreement between the parties (i.e., the beneficiary or the victim does not wish for the benefit or damage and whoever causes the damage or benefit does not pay for it). The LCC structure proposed by the ISO can be summarised, based on the LCA model, in four stages: 1. 2. 3. 4.

Definition of the objectives and reference framework Definition of the cost categories per life cycle Calculation of the cost variables and analysis of future costs Interpretation of the results.

According to the specifications in CEN/TC 350, the sustainable performance of a building in Europe must consider the three environmental, economic and societal dimensions based on a common protocol (cfr. Chap. 1). The three performances can be considered separately, but based on a single functional equivalent that takes into account the client’s needs and the functional and technological requirements which must be met, e.g., structural safety, fire safety, accessibility, disassemblability, recyclability, maintainability, durability, and the service life of a building or part of the construction. EN 15643-4 Sustainability of construction works—Assessment of buildings—Part 4: Framework for the assessment of economic performance, and EN 16627—Sustainability of construction works—Assessment of economic performance of buildings—Calculation methods are the standards defining the method to calculate economic performance in line with the general sustainable performance

48

3 Life Cycle Methodologies

framework for the construction sector (EN 15643-1, see Fig. 1.7 in Chap. 1). They are complementary to the international standards regarding sustainability assessment (ISO 15392) and the ISO 15686 series. Economic performance assessment considers the building and its surroundings and uses the informative modules in the system of standards developed by CEN/TC 350, dividing the costs and, if relevant, the revenue, in the following modules: A0 (pre-construction stage), A1-3 (product stage), A4-5 (construction stage), B (use stage), C (end-of-life stage) e D (benefit and loads beyond system boundary).

3.2.1.2

The Code of Practice of Environmental LCC: SETAC WG5

Based on the analysis of the results of several European LCC studies, SETAC defined three types of Life Cycle Costing (Hunkeler et al. 2008; Swarr et al. 2011): Conventional LCC: assessment method taking into account all the costs associated with the life cycle of a product that are directly sustained by the main producer or final user. Assessment focuses on real internal costs, sometimes without even considering use and end-of-life costs if the latter are sustained by others. A conventional LCC is not generally accompanied by results from a life cycle assessment (LCA). The perspective is chiefly that of a market actor: the producer or user/consumer (Fig. 3.3). Environmental LCC: this assessment method considers all the costs associated with the life cycle of a product that are directly sustained by one or more actors involved in the whole life cycle (supplier, producer, user/consumer and person responsible for end of life), including the externalities which are envisaged to be internalised in future decision-making processes (Rebitzer and Hunkeler 2003). In other words, the Environmental LCC fine-tunes the Conventional LCC; on the one hand it requires the inclusion of all the life cycle stages and future costs to be sustained and, on the other, it takes into account the environmental costs calculated using the LCA method. This data is kept separate and not monetised. The perspective is that of one or more actors of a given market. If relevant, the environmental LCC analysis can include incentives and taxes. Societal LCC: this assessment method considers all the costs associated with the life cycle of a product that are directly sustained by anyone in society, now and in the

Fig. 3.3 A producer’s perspective in a conventional LCC (by Hunkeler et al. 2008)

3.2 Life Cycle Costing

49

long-term future. Societal LCC includes all the Environmental LCC plus additional assessments of further external costs, usually in monetary terms (e.g., based on the willingness-to-pay method). The perspective is that of the whole of society, internationally and nationally, including governments. Compared to the Environmental LCC, incentives and taxes do not affect the net assessment of costs and are therefore not included in the Societal LCC. SETAC proposes to find a solution in order to overcome the limits of the Conventional LCC, i.e., to not only broaden the perspective by including all the actors involved in the life cycle, but also consider all the life cycle stages. Actors play a key role, as does the viewpoint from which the LCC analysis is performed. Figure 3.3 shows how the perspective of a conventional analysis changes depending on who performs it. The example illustrates an analysis performed by a producer of building components. Furthermore, when considering all the life cycle stages it’s important to distinguish between LCC Planning and LCC Analysis. The former is implemented during design to compare several alternatives while the latter uses monitoring analysis to evaluate the performances of products or existing buildings. Environmental LCC is not a stand-alone technique with which to monetise externalities (e.g., environmental impacts), but is considered complementary to the LCA analysis. The work by SETAC therefore falls within the broader concept of Life Cycle Sustainability Assessment (LCSA) sponsored by the United Nations Environmental Programme (UNEP) for the application of sustainability to life cycle. According to the LCSA principles, as a stand-alone analysis, the LCA for the environmental life cycle assessment, the LCC for the environmental life cycle costing and the SLCA for the social life cycle assessment, have to be applied in parallel, assuming the same system boundaries and the same functional unit (Klöpffer 2003; Swarr et al. 2011). Although the work by SETAC applies the LCC concepts to a number of commodity sectors, it is in line with the sustainability standards developed in the construction sector examined earlier.

3.2.2 General Framework and Costs Like the LCA standards, four LCC stages were identified based on the results of the analysis of the environmental LCC model and reference standards (Notarnicola et al. 2005): 1. 2. 3. 4.

Goal and scope definition Cost categories (economic life cycle inventory) Cost analysis and discounting Interpretation, reporting, and review.

50

3 Life Cycle Methodologies

Fig. 3.4 LCC diagram based on ISO 15686-5 (Source by the author)

Figure 3.4 shows the framework developed to apply the LCC to design based on the LCDM model illustrated in Chap. 4. 1. Goal and scope definition The objective of a life cycle costing analysis is to quantify the costs to be calculated in a decision-making or assessment process that must consider the client’s brief and the requirements of several analytical fields (e.g., assessments regarding the environment, design, safety, functionality, and regulations). Quantification must be sufficiently detailed depending on the design stage (concept design, developed design and technical design). In addition, the reference framework must be decided together with the client so as to clarify the system boundaries (costs included/ excluded by the LCC analysis). The LCC analysis must envisage a pre-established list of costs to be sustained during the physical, technical, economic, and functional life of a building, component, or product, considering the prefixed period of analysis. If the analysis is part of special procurement contract strategies, or its objective is to promote sustainability, then it’s possible to include both the costs sustained by several actors and non-construction costs (e.g., cost of land), taxes and incentives, revenue from the asset, and externalities. The complete analysis will therefore be called Whole Life Cycle (WLC) (ISO 15686-5). The designer and the client will define the boundary of the system, schematising the cost items to be included/ excluded from the analysis and dividing them according to the structure in Fig. 3.5 throughout the life cycle.

3.2 Life Cycle Costing

51

Fig. 3.5 LCC and WLC system boundaries based on ISO 15686-5

2. Cost categories (economic life cycle inventory) After establishing the objectives and the boundaries of the analytical system, it is now time to define the cost variables of the life cycle. This procedure could be called the LCC “inventory stage”. It includes: • Acquisition costs • Use costs due to: – – – – –

operations (energy consumption, water consumption, etc.) maintenance (preventive, corrective, urgent) management of maintenance activities (inspections, design, etc.) cleaning and ordinary maintenance interventions indirect maintenance costs (loss of use, closure, etc.)

• End-of-life costs • External costs/ revenue (taxes and incentives, revenue, externalities, other costs/benefits). According to the Italian regulations governing public procurement contracts, life cycle costs must also include “costs attributable to environmental externalities associated with products, services or works during the life cycle, so long as their monetary value can be determined and verified. Said costs can include the costs of greenhouse gas emissions and other polluting substances, as well as other costs associated with the easing of climate change”.20 An environmental externality, e.g., pollution, does not directly affect the economic conditions of the user of the product/service/work, but is borne by the local, national or international community. 20 DGLS

50/2016, art. 96, 1b.

52

3 Life Cycle Methodologies

The very few methods that exist to monetise externalities are difficult to apply due to the geographical and temporal limits of the data. Nevertheless, a case study presented in Chap. 4 of this book tests the possible monetisation of the environmental impact.21 The European Directive 2009/33/EC also provide a method to promote.22 In Europe there are several projects aimed at monetise environmental costs. For instance, the project ExternE, “Externalities of Energy”, launched in 1991 is the result of more than twenty research projects conducted till 2005.23 3. Cost analysis and discounting As emphasised earlier, the purpose of a Life Cycle Costing (LCC) analysis is to quantify the life cycle costs associated with an evaluation or decision-making process including, where possible, inputs from other evaluations (e.g., environmental assessments, design assessments, functional assessments, regulatory assessments, etc.).24 Therefore, the objective of the classification stage (which could correspond to the “impact assessment” of a LCA) is to use impact indicators to quantify the costs to be sustained in each of the life cycle stages. The costs of the LCC analysis can be expressed in “real costs” or “nominal costs”25 and can be sustained either now or in the future, actualised or not, through discounting. Whatever the case may be, the method to be used must first be stated within the boundaries of the LCC analysis system. Standards and regulations suggest using “real costs” and “discounting” future costs. The “discounting” mechanism is the financial process that determines the current value of a capital, given a specific future date of return. By applying a discount, it is possible to identify a financial equivalent between two capitals with different return dates. More specifically, it involves introducing the concept of the “time value of money”, i.e., the value of money at the exact moment when it will be received or paid. The time value of money consists in discounting future costs to reflect the loss of value that takes place in the transition year compared to the current reference year. In an LCC assessment it is important to establish in which year each item will be sustained. 21 Cfr.

Sect. 4.4. Art. 6 of Directive 2009/33/EC of 23 April 2009 “Methodology for the calculation of operational lifetime costs” and Table 2 attached to the costs of emissions in road transport for the year 2007: CO2  0.03–0.04 e/kg; NOX  0.0044 e/g; NMHC  0.001 e/g; particulate matter  0.087 e/g. 23 The main scope of the first ExternE project has been the airborne pollutants from power plants and the development of the Impact Pathway Analysis (IPA). Then, main goals of the follow-up projects have been on the one hand improving and extending the methodology and incorporating new knowledge, on the other hand extending the field of applications, such as heat production, transport, and industrial activities (http://www.externe.info, access on 15/01/2019). 24 In the construction sector, the calculation method and economic impact indicators can be assumed by ISO 15686-5, EN 16627 and EN 15643-4. 25 Real costs correspond to the current value, while nominal costs are obtained by multiplying real costs by the inflation/deflation rate linked to the percentage of increase/decrease in prices per year, from the initial reference date to the year in which the cost will be sustained. 22 Cfr.

3.2 Life Cycle Costing

53

Discounting is calculated based on future costs and reducing it by a qd factor derived from discount rate d using the following equation: qd 

1 (1 + d)n

where: d discount rate envisaged per annum n number of years between the current reference year and the year in which the cost will be sustained. This equation allows real costs to be converted into discounting costs using the qd , the factor that will permit both the correlation of present and future monetary values and comparison of costs without considering inflation (which is uncertain, especially over long periods of time). The Net Present Value (NPV) can then be calculated. The NPV is the discounting of the economic flow (positive flows—benefit—less negative flows—costs) using the following formula: N PV 

p    (Bn × q) − (Cn × q)  n1

Bn Cn − n (1 + d) (1 + d)n



where: Cn Bn q d n p

cost in year n benefit in year n discounting factor discount rate envisaged per annum number of years between the current reference year and the year in which costs are incurred period of analysis.

It’s important to choose the right discount rate d because variations in its value cause variations in the magnitude of future costs. The lesser the discount rate, the greater the influence of future costs in the LCC analysis and vice versa. Accordingly, using higher discount rates tends to benefit solutions with lesser initial costs. Standard 15686-5 establishes a range of values from 1 to 6%. To maximise comparability of results, EN 16627 adopts a 3% discount rate26 which can, however, be integrated using supplementary calculations with different discount values. 4. Interpretation and risk analysis Since the LCC methodology requires theories to be formulated regarding future behaviour, the standard 15686-5 suggests a risk analysis be performed to gradually 26 The

3% discount rate was taken from Delegated Regulation (EU) n. 244/2012 for the calculation of the Cost Optimal according to Directive 2010/31/EU.

54

3 Life Cycle Methodologies

reduce uncertainty, despite the fact a residual risk continues to remain. The “risk” concept differs from that of “uncertainty”: the former is used to indicate measurable probability, the latter indicates a probability to which it is impossible to associate a value. A sensitivity analysis can be performed to establish how variations in the uncertainty range can influence the correctness of the evaluated and compared options. There is one condition: that the value range be probable, within the limits of the boundaries of the system, and in line with the requirements contained in the client brief. These analyses can help identify which of the input data taken into consideration cause the greatest impact on LCC results and how reliable the final result will be. Certain variables will have a greater effect on the degree of uncertainty, for example the discount rate, the period of analysis, incomplete data (costs based on hypotheses, incorrect service life estimates, etc.), or non-implementable activities (maintenance, restoration or replacement cycles). To verify the results, standards suggest documenting the analysis using a series of summaries, including the definition of the scope, reference framework, initial hypotheses, limits of the boundaries, restrictions, uncertainties, risk, and effects. This will help any user to easily control the results and ensuing products.

3.3 Combining LCA and LCC “Traditional” Life Cycle Costing is a method to evaluate the most advantageous investment between alternative scenarios. Nevertheless, despite the fact that the system boundaries include a life cycle perspective that also considers operational costs and the initial investment, environmental costs are not included in LCC. In fact, it is important to emphasise that “traditional” LCC does not turn into an environmental calculation model just because it contains the words life cycle (Braume 2006). The multiple definitions of Life Cycle Costing often refer to global environmental costs (Whole Life Cost, Full LCC, Life Cycle Cost Assessment, Environmental LCC). Baume summarises the differences between these multiple meanings contained in the literature of corporate environmental accounting tools.27 The combination of terms such as full, total, true, whole with life cycle and environmental gives these cost assessment models a sort of “environmental aura” conveying the efforts to promote integration between the two concepts: environmental impact and costs (Braume 2006). Since the nineties many studies and researches have focused on the possibility of integrating environmental aspects into Life Cycle Costing (Settanni et al. 2012).

27 Corporate

environmental accounting processes are the identification, measurement, calculation, analytical, preparatory, interpretation, and communication processes used in the management of financial (and non-financial) data to plan, assess and control the environmental aspects of a company. (Van der Veen 2000, pp. 155–175).

3.3 Combining LCA and LCC

55

In particular, scholars have asked themselves whether merging these two aspects is the correct line of action and which approach should be adopted. According to ISO 14040/44, life cycle approach and methodologies described in that International Standard can be applied also to economic or social aspects of a product. Likewise, the standards developed by CEN/TC 350 define the overall sustainability performance of a building by using an assessment that considers in parallel the economic, environmental, and societal aspects of the life cycle. As mentioned earlier, the Environmental LCC proposed by SETAC intends to fine-tune the conventional LCC by requiring not only the inclusion of all the life cycle stages and future costs to be sustained (by all the actors involved), but also a separate calculation for environmental effects (externalities that cannot be monetised). These effects are assessed using a LCA analysis in line with ISO 14040-44 (2006). Social aspects are also included in Societal LCC. In the Life Cycle Sustainable Assessment (LCSA) proposed by the UNEP (Valdivia et al. 2011), the three dimensions of sustainability are analysed separately based on the same goals and scope assumptions. Likewise, and despite the fact that the objective of ISO 15686-5 is to establish the rules governing the application of a traditional LCC analysis based on an envisaged service life of buildings and their components, the standard does however try to achieve integration between the aspects associated with sustainability, as advocated by the Environmental LCC model proposed by SETAC. In particular as regards two elements. The first is the introduction of the Whole Life Costing concept (WLC) that broadens the boundaries of the LCC analysis to include (“where possible”) externalities, environmental impact costs, and societal costs and benefits.28 The second is a direct reference to sustainability,29 in particular, the intention to merge the economic analysis with ISO 14040/44 (LCA) in order to measure the impact of environmental externalities. Apart from the aforementioned projects, several studies are also presented to complete the reference framework. The studies are scientifically interesting insofar as they are preceded by and sometimes assist in the drafting of the standards. They were developed to disseminate Life Cycle Assessment (LCA) and Life Cycle Costing (LCC) within the field of construction. The points of view are heterogeneous because the studies were performed by actors from different sectors, for example: international organisations for the development of norms and standards (ISO, CEN, ASTM), international scientific communities (SETAC), public authorities (City of Hong Kong, government agencies (NIST, EMSD) and professional studios (Arup, Davis Langdon). The following list contains the records that focus on the relationship between the economy and the environment and provide a theoretical-methodological contribution that helps to identify the specific objectives of the model and methodology adopted in this book.30

28 ISO/FDIS

15686-5 paragraphs 6.2, 6.3, 6.4. paragraphs 6.5. 30 Cfr. Chap. 4 (Life Cycle Design Model—LCDM) 29 Ibidem,

56

3 Life Cycle Methodologies

• Building for Environmental and Economic Sustainability (BEES) (Lippiatt and Boyle 2001; Lippiatt 1998; Lippiatt and Means 2014): a tool designed by the National Institute of Standards and Tecnology (NIST) based on ASTM Standard E917-99 “Standard practice for measuring life-cycle costs of buildings and building systems” and ISO 14044 respectively for the economic and environmental assessment of building products. • Davis Langdon Management Consulting: “Life cycle costing (LCC) as a contribution to sustainable construction: a common methodology”. May 2007. • Project by EMSD and Arup: “Life Cycle Assessment (LCA) and Life Cycle Cost (LCC), Tool for Commercial Building Developments in Hong Kong”. June 2006. • European Commission. Task Group 4: “Life Cycle Costs in Construction. Final Report” (July 2003). All the aforementioned studies tend to identify a set of common rules used to develop the environmental and economic life cycle analyses. Some try to combine the results in a single numerical or graphic assessment, giving the design team and the client the possibility to assign higher priority to: economic and environmental aspects; the initial stage or use stage; energy consumption during the operational stage or the material product stage. The theoretical integration of the separate results of the analyses into a final output provides dual benefits for the dissemination of a life cycle approach to the project. On the one hand, separate application of the methods ensures transparency of the data and the individual results of the economic and environmental analyses. On the other, final synthesis in a single output increases the ability to communicate and exchange information at the decision-making level. In fact, if the goal of the analysis, as suggested by the standard, is to satisfy the requirements of the client brief, then the results have to also be accessible to non-expert users of LCA and LCC. Based on these considerations, the life cycle stages to be included in the development of a common method will now be analysed.

3.3.1 A Common Framework for the Building Sector The growing importance of the need to include environmental and economic impacts in all the design stages emerges from the analysis of available studies and standardisation tools developed to comply with European Directives and new requirements. Moreover, designers are increasingly asked to focus more on an integrated design incorporating post-construction life cycle scenarios in terms of costs, energy consumption, and pollution. Sustainability tools can be included in the design stage; these tools should be based on the life cycle approach according to the objectives defined during the various design levels. Starting with the initial design stages, design strategies are established during the conceptual and schematic design stage. The typologies of building components and systems are chosen during the ensuing design development stage;

3.3 Combining LCA and LCC

57

Fig. 3.6 Relationship between requirements, design levels, and life cycle stages of the building

these components and systems have to satisfy the requirements identified during the initial design stage and specified in the client brief. Finally, products, installation techniques, and everything else required to construct the building are selected in the construction documents stage. Each design level contains increasingly complex and complete data used to calculate performance. The control tool needed to manage this process must take into account the design parameters and increasing levels of definition of the project, bearing in mind the effects of design on the life cycle stages (Fig. 3.6). This tool can be a useful instrument to share the results during the design stages, both between the designers, and also between the designers and client. The “design matrix”, presented in the next chapter, is a design tool based on this methodological reference framework.

3.3.1.1

Building Life Cycles

The life cycle of a building, or of one of its components, is linked to its service life. Standard ISO 15686-1 defines the principles regarding its envisaged service life. Service life planning is “a planning process capable of ensuring that the service life of a building, or other architectural asset, be equal to or greater than its design life”. While “design life” means the envisaged life formulated by the designer, “service life” is the period of time during which the building reaches or exceeds its initial required performance. The standard also specifies that “if required, the envisaged service life planning can take into account the cost of the life cycle of the building and its impact on the environment”.

58

3 Life Cycle Methodologies

The above explains the important difference between “service life” and “service life of the component parts”. The service life of a building is decided by the client, owner, or manager, while the service life of its components is extrapolated from the individual technical and functional characteristics of the parts themselves, their position in the building, or their obsolescence (Molinari 2002). Three specific forms of the service life of a building can be identified based on these premises: • physical service life: corresponding to the lifespan enabled by physical deterioration processes, compared to the initial performance; • functional service life: corresponding to the lifespan enabled by the “physical service life” and obsolescence compared to the performance/requirements ratio; • economic service life: corresponding to the capacity of the building to ensure an income and have an economic value. When performing a LCA or LCC study it’s important to establish the “period of analysis”. This period corresponds to the duration of the study, but may not necessarily coincide with the service life of the building. For example, as regards cost analysis it corresponds to the “economic service life” of the building which obligatorily presupposes conservation of the physical and functional service life. The economic service life can end even if the other two do not. This means that the end of life moment does not necessarily coincide with the end of the physical or functional service life of the building, but that it occurs at the end of the study period established by the client’s brief (the plan drafted by the client or owner). Furthermore, the period of analysis of the Life Cycle Assessment model often differs from that of a cost assessment model. Figure 3.7 indicates the life cycle stages considered for some of the projects that were analysed (LCA in green and LCC in blue). In most of the reference projects the environmental assessment period begins with the acquisition of raw materials and terminates with the end of life. Instead the economic assessment period begins when payment is made and future investments are planned (i.e., when building products are purchased and used) and ends at a point that does not necessarily coincide with the end of life stage.31 In some cases (Environmental LCC and WLC), it’s useful to consider the design and research costs as well as the pre-construction stages. Product and construction process stage As shown in Fig. 3.8, the two methods, LCA and LCC, differ the most in the initial stage which includes the following processes32 : (a) Acquisition/extraction of the raw materials (b) Transportation and processing of the raw materials 31 There are cases in which the economic investment ends before the decommissioning stage or cases in which the actors involved change, modifying the goals and scope of the investment, and hence the period of analysis. 32 This list is indicative and can be implemented with other specific processes depending on the system boundaries of the project to be assessed. For example, it can include information about the site or design activities, site management, temporary worksite structures, etc.

3.3 Combining LCA and LCC

59

Fig. 3.7 Life cycle analysis in the reference projects. Comparison of LCA (green) and LCC (blue) life cycle stages

Fig. 3.8 Stages of the assessment model and indicators. LCA/LCC comparison

(c) (d) (e) (f) (g) (h) (i) (j)

Disposal of waste from (b) Production of products/building elements Disposal of waste from (d) Research and design (R&D) Acquisition of products/building elements Transportation of products/building elements to the worksite Installation of the products/building elements Disposal of waste from (i).

Each process can require the involvement of one or more actors who interact with the system. Apart from the difficulties inherent in obtaining the information, in a LCA analysis it would be hypothetically possible to recreate the entire production chain

60

3 Life Cycle Methodologies

of an element, working backwards to the extraction of the raw materials. Instead in economic terms this would be impossible bearing in mind that: • The economic analysis period begins with the investment: from the acquisition of the products (g) or, possibly, from the design process (identifying the previous research process is realistically almost impossible). • The cost of purchasing the building products (supply cost) coincides with the market price (costs + marginal revenue).33 • The price of raw materials varies more rapidly than the price of finished products.34 • Persuading companies to provide disaggregated costs would not be credible.35 First of all because it would almost certainly not involve a single actor but multiple actors (producer, machinery supplier, company extracting raw materials, transportation, etc.). Secondly, because a product is often part of a chain that includes other products. It would be very difficult to separate the costs and associate them with a single product. • The fact many actors are involved and there are multiple market transactions could cause errors in terms of double-counting. The proposed model (see Chap. 4) was inspired by the Environmental LCC model that suggests the use of market prices, except when the assessment is performed within the production company itself. Only in this case could it be possible to transparently use disaggregated cost data. Summing up, the analysis of the reference methods and projects have shown that two options are available in LCC assessments. The first is to consider the costs of the initial stage from the company viewpoint and divide them into the cost of raw materials, labour, capital, energy, inflation, profits, R&D, and external costs. The second option is to consider the market price. The result is identical. Problems could arise in the second, simpler option if it were impossible to obtain the prices of technologies not yet on the market, or if the sensitivity analysis took into account the variables within the company (Huppes and Ciroth 2008). Use stage The use stage lasts the longest and therefore plays a very important role in the global assessment of a building. It is also the most aleatory because it is based on future scenarios that are difficult to foresee and is subject to high levels of uncertainty. Two categories exist: operation and maintenance. The operational stage generally involves energy and fuel demand (Operational Energy), water use and disposal, waste management, greenhouse gas emissions 33 The market price of a product is the result of the difference between revenue and costs (raw materials, labour, capital). 34 The revenue margin (regarding market prices) considers the possible variations of production costs (including fluctuation of the costs of raw materials). If price variations do not exceed the envisaged variation, the market price remains the same. 35 One reason for a slowing down of the dissemination of the standardised global cost assessment amongst companies was, in many cases, shown to be due to non-correspondence to the cost assessment system used by individual companies or organisations.

3.3 Combining LCA and LCC

61

(Operational Carbon), and the safety of the building. The functioning of the whole building is closely linked to its use, location, type of system, type of user, etc. The operational stage must also take into account scenarios that envisage the possible loss of the “physical duration” of a material or component (and consequently the use of the building) if this element cannot be restored either by maintaining or replacing it. In this case the assessment should include the quota linked to the increase in energy required during use to maintain the initial performance conditions.36 Maintenance also has to be considered during the use stage. Over a period of time all buildings are inevitably subject to deterioration and functional, technological, and economic obsolescence. The maintenance stage includes the activities required to maintain the functioning of the asset and system network according to the established conditions. For example, the envelope requires cyclical maintenance (painting, renovation, repair) and the replacement of materials or components. Although envelope maintenance is often ignored by LCA or LCC studies, it becomes crucial when a life cycle assessment is performed during the design stage, i.e., when choices are made regarding the building technology that has to be adopted to reduce the consumption of energy required to heat the building in the winter and cool it in the summer. When an element is either replaced or regenerated, the disposal of the replaced or removed element must be considered together with the new envisaged works. Despite the fact that scientific documentation regarding the duration of materials is either scarce or difficult to find, it is important to theorise a durability value for each material based on the information in handbooks (often based on American or British data),37 producers’ guarantees (frequently found in insurance documents) or even historical data (at least for traditional technologies).38 This makes it possible to identify, from the outset, a number of cyclical maintenance interventions and/or replacements throughout the life of the building. These interventions will have to be considered in the assessment as additional embodied energy, environmental impacts, and operational costs. Less important environmental and economic interventions, e.g., cleaning and ordinary maintenance, can be included if they are relevant to the objectives of the study. End-of-life stage The end-of-life stage of the theorised period of analysis represents the end of the “service” life of a building, i.e., the moment when a decision is taken to refurbish or deconstruct the building. The choice depends on the type and function of the building, on its degree of obsolescence (which can theoretically equal the end of the estimated end-of-life period), on certain technological choices, or on other causes that are difficult to predict during the design and construction stage. As highlighted by the review of the experiences and studies on integration between LCA and LCC, the end-of-life assessment is often considered optional. In actual fact 36 Cfr.

Appendix A of Chap. 4. Molinari (2002), BEES (2006), ASTM 917; ISO 15686, Manfron (1998), Perret (1995). 38 It is impossible to know the functional duration of certain products on sale since they have only just come onto the market. 37 Cfr.

62

3 Life Cycle Methodologies

it is very important because it can modify—even radically modify—the technological choices made during the design stage. Unfortunately the data regarding both environmental and economic assessments are extremely aleatory. This is because the envisaged hypothetical scenarios take place many years after they are theorised. In fact, future restoration, disposal, and recycling technologies may be very different. Approximation also affects costs. Direct contact with the companies involved with waste recovery and disposal revealed that costs are extremely uncertain, even from one trimester to another.39 During the design stage it may be useful to theorise one or more end-of-life scenarios so that design focuses more on sustainability (i.e., design for disassembly approach) making it always possible to update the information in the analysis during the useful life of the building.

References Baldo GL, Marino M, Rossi S (2005) Analisi del ciclo di vita LCA. Materiali, prodotti, processi, Edizioni Ambiente, Milano Blengini G, Di Carlo T (2010) The changing role of life cycle phases, subsystems and materials in the LCA of low energy buildings. Energy Build 42:869–880 Bonanomi M, De Flumeri C, Lavagna M (2012) Edifici a consumo energetico zero. Orientamenti normativi, criteri progettuali ed esempi di Zero Energy e Zero Emission Buildings, Maggioli Boustead I, Hancock G (1979) Handbook of industrial energy analysis. The Open University, Milton Keynes, Hellis Horwood Limited, Chichester, West Sussex, England Braume A (2006) Linking life cycle costing and LCA for building and construction—a framework for eco-efficient construction. In: Proceedings, 2nd international conference on quantified ecoefficiency analysis for sustainability, EgmondaanZee, 28–30 June 2006 Davis Langdon Management Consulting. Life Cycle Costing (LCC) as a Contribution to Sustainable Construction: A Common Methodology—Final Methodology. 2007. http://ec.europa.eu/ enterprise/sectors/construction/studies/life-cycle-costing_en.htm. Accessed 16 Sept 2018 Deng S, Wang RZ, Dai YJ (2014) How to evaluate performance of net zero energy building—A literature research. Energy 71:1–16 EMSD, ARUP, Life Cycle Assessment (LCA) and Life Cycle Costing (LCC) Tool for Commercial Building Developments in Hong Kong—User Manual (2005) EN 15643-4:2012 Sustainability of construction works—assessment of buildings—part 4: framework for the assessment of economic performance EN 16627:2015—Sustainability of construction works—assessment of economic performance of buildings—calculation methods EN 16627:2015 Sustainability assessment of buildings—economic performance of buildings Estevan H, Schaefer B, Adell A (2018) Life cycle costing state of the art report European Commission—Joint Research Centre (2010) Institute for environment and sustainability: international reference life cycle data system (ILCD) handbook—general guide for life cycle assessment—detailed guidance. Publications Office of the European Union, Luxembourg European Commission Task Group 4: Life Cycle Costs in Construction. Final Report, June 2003 Fregonara E (2015) Valutazione sostenibilità progetto. Life Cycle Thinking e indirizzi internazionali. Franco Angeli, Milano

39 The

information was provided during interviews with operators in the Piedmont Region (Italy).

References

63

Fregonara E (2017) Methodologies for supporting sustainability in energy and buildings. The contribution of project economic evaluation. Energy Proc 111:2–11 Giordano R (2010) I Prodotti per l’edilizia sostenibile.La compatibilità ambientale dei materiali nel processo edilizio, SistemiEditoriali, Napoli Giordano R, Serra V, Tortalla E, Valentini V, Aghemo C (2015) Embodied energy and operational energy assessment in the framework of nearly zero energy building and building energy rating. Energy Proc 69 Glisoni M, Bianco E, De Leonardis D (2010) Progetto APE—Linee guida per l’applicabilità della metodologia Life Cycle Costing agli appalti pubblici ecologici. ARPA Piemonte Gluch P, Baumann H (2004) The life-cycle costing (LCC) approach: a conceptual discussion of its usefulness for environmental decision-making. Build Environ 39:571–580 Guinée JB, Gorrée M, Heijungs R, Huppes G, Kleijn R, de Koning A, Van Oers L, Wegener Sleeswijk A, Suh S,. Udo de Haes HA, De Bruijn JA, Van Duin R, Huijbregts MAJ (ed) (2002) Handbook on life cycle assessment: operational guide to the ISO standards. Series: Eco-efficiency in industry and science. Kluwer Academic Publishers. Dordrecht Hammond GP, Jones CI (2008) Embodied energy and carbon in construction materials. Proc Inst Civil Eng Energy 161(2):87–98. https://doi.org/10.1680/ener.2008.161.2.87 Hunkeler D, Rebitzer G, Lichtenvort K (ed) (2008) Environmental life cycle costing. CRC Press, New York Huppes G, Ciroth A, Lichtenvort K, Rebitzer G, Schmith W, Seuring S (2008) Modelling for life cycle costing. In: Hunkeler D, Rebitzer G, Lichtenvort K (eds) Environmental life cycle costing. CRC Press, New York IEA (2014) Energy in buildings and communities programme. In: Annex 57: evaluation of embodied energy and CO2 eq for building construction ISO 14040:2006 Environmental management—life cycle assessment—principles and framework ISO 14044:2006 Environmental management—life cycle assessment—requirements and guidelines ISO 15686-5:2017 Building and constructed assets—service life planning—life cycle costing ISO 21929-1:2011 Sustainability in building construction—sustainability indicators—part 1: framework for the development of indicators and a core set of indicators for buildings Karsten V, Riley M (2009) IEA joint project: towards net zero energy solar buildings (NZEBs), IEA SHC Task 40—ECBCS Annex 52 Klöpffer W (2003) Life-cycle based methods for sustainable product development. Int J Life Cycle Ass 11(special issue 1):116–122 Lavagna M (2008) Life cycle assessment in edilizia. Hoepli, Milano Lippiatt BC (1998) BEES: balancing environmental and economic performance. Construct Spec 51:4 Lippiatt BC, Boyle AS (2001) Using BEES to select cost-effective green products. Int J Life Cycle Ass 6:76–80 Lippiatt BC, Means RS (2014) Evaluating products over their life cycle, green building, 3rd edn. Wiley, Hoboken Manfron V (1998) La manutenzione delle costruzioni. In: Manfron V, Siviero E (eds) Manutenzione delle costruzioni. Utet, Torino Molinari C (2002) La manutenzione come requisito di progetto. Sistemi Editoriali, Napoli Monticelli C, Thiébat F (2016) Energy and environmental performances assessment/Valutazione delle prestazioni energetico-ambientali. In: Lucarelli MT, Mussinelli E, Trombetta C (ed) Cluster in progress, pp 109–113. Maggioli, Santarcangelo di Romagna (RN) NIST (2000) BEES 2.0: building for environmental and economic sustainabiliry technical manual and user guide. NISTIR 6520. National Institute of Standards and Technology (NIST). Gaithersburg, MD, USA Notarnicola B, Tassielli G, Settanni E (2005) Life Cycle Costing nella produzione dell’energia elettrica, lineamenti metodologici e applicazione. Ambiente, Risorse, Salute (ARS), n. 101 Perret J (1995) Guide de la maintenance des bâtiments. Le Moniteur, Paris Pollo R (2015) Progettare l’ambiente urbano. Riflessioni e strumenti. Carocci, Roma

64

3 Life Cycle Methodologies

Publisher: ICLEI—Local Governments for Sustainability, European Secretariat Rebitzer G, Hunkeler D (2003) Life-Cycle Costing in LCM: ambitions, opportunities and limitations. Discussing a framework. Int J Life Cycle Ass 8:253–256 Rebitzer G, Nakamura S (2008) Environmental life cycle costing. In: Hunkeler D, Rebitzer G, Lichtenvort K (ed) Environmental life cycle costing. CRC Press, New York Saur K, Donato G, Cobas Flores E, Frankl P, Jensen AA, Kituyi E, Lee KM, Swarr T, Tawfic M, Tukker A UNEP/SETAC life cycle initiative, draft final report of the LCM definition study, version 3.6, 17 Nov 2003. https://www.lifecycleinitiative.org/wp-content/uploads/2012/12/2003%20-% 20LCM%20Definition%20Study%20-%20v3.6.pdf. Accessed 02 Nov 2018 Settanni E, Notarnicola B, Tassielli G (2012) Life cycle costing (LCC). In Fogel D, Fredericks S, Spellerberg I (eds) The encyclopedia of sustainability, Vol 6: measurements, indicators, and research methods for sustainability, pp. 221–224. Berkshire Publishing, Great Barrington, MA Swarr ET, Hunkeler D, Klopffer ¨ W, Pesones H, Ciroth A, Brent AC, Pagan R (2011) Environmental life cycle costing: a code of practice. SETAC, New York The Intergovernmental Panel on Climate Change (IPCC) Assessment Reports. http://www.ipcc.ch. Accessed 5 Dec 2018 Thormark C (2002) A low energy building in a life cycle—its embodied energy, energy need for operation and recycling potential. Build Environ 37:429–435 Valdivia S, Ugaya CML, Sonnemann G, Hildenbrand J (2011) Towards a life cycle sustainability assessment. Making informed choices on products. UNEP, Paris Van der Veen M (2000) Environmental management accounting. In: Kolk A (ed) Economics of environmental management, Harlow, UK, Pearson Education Ltd Voss K, Musall E (2012) Net zero energy buildings: international comparison of carbon-neutral lifestyles. Birkhäuser, München Wolf MA, Pant R, Chomkhamsri K, Sala S, Pennington D (2012) International reference life cycle data system (ILCD) handbook—towards more sustainable production and consumption for a resource-efficient Europe. JRC Reference Report, EUR 24982 EN. European Commission—Joint Research Centre. Luxembourg. Publications Office of the European Union Wright LA, Kemp S, Williams I (2011) ‘Carbon footprinting’: towards a universally accepted definition. Carbon Manag 2(1):61–72

Chapter 4

Defining an Innovative Design Method Based on the Life Cycle Approach

Abstract This chapter will illustrate an experimental design model that can be used to produce two different outcomes based on two levels of assessments. This tool is called the Life Cycle Design Model (LCDM).

The first level of assessment involves creating a grid that is useful in the design process; the grid includes a number of design parameters such as, amongst others, the costs and environmental/energy impacts of a building or building component. In particular, design concepts and project values are shown in a matrix for each life cycle stage (concepts or parametric values for the early design phase, and numbers for the schematic and detailed design phases). The design matrix allows designers to not only benchmark their project values against those mandatory by law, but also compare several solutions. The first part of this chapter illustrates how to build the matrix. The second level of assessment, illustrated in the second part of this chapter, consists in exploiting LCA and LCC methods to develop a user-friendly tool for designers and other actors involved in the building process so that they can assess the most sustainable design option. Using points and weights the model expresses an economic-environmental efficiency value (eCO) that is more subjective and controversial than the design matrix, but still useful as a tool to steer design processes based on the initial architectural concepts. The LCDM model, using the design matrix and the eCO factor, is also helpful when talking to non-expert clients and communities.

4.1 Design with a Matrix A building should be designed to last for a pre-established lifetime; the perspectives of the actors involved (planners, designer, users, society, environment, etc.) should be part of this process. Decision-making methods should be used in the early stages of the life cycle design process to define the expected life of the building and check the sustainability of the solutions adopted at each design level: these solutions regard © Springer Nature Switzerland AG 2019 F. Thiébat, Life Cycle Design, PoliTO Springer Series, https://doi.org/10.1007/978-3-030-11497-8_4

65

66

4 Defining an Innovative Design Method …

Fig. 4.1 Integrated design: concept map

strategy, the system adopted, and details. The decisions taken during the early stages of the design process will have greater impact on in use performance and final costs (Østergård et al. 2016). The Life Cycle perspective requires stakeholders to find new ways to cooperate. So, integrated design provides a holistic approach and ongoing consideration of numerous aspects. Here, the interoperability of design tools is a key factor; it supports the design team in their assessment of the life cycle of buildings, the cost of that life cycle, energy performance, and all the design process parameters such as social aspects, aesthetic value and urban connection. All the parameters can be integrated in a matrix. What is the meaning of “designing with a matrix”? What is the relationship between the matrix and the life cycle in an integrated design approach? Who are the stakeholders involved in the matrix? How it can be applied? Figure 4.1 illustrates the integrated design approach. Links are created within the boundary of sustainability between the final goal of the project (‘WHAT’) and the actors involved (‘WHO’). In accordance with the economic definition of sustainability, their common goal is to ensure well-being and quality while maintaining services and natural resources over a period of time (Pearce and Turner 1991). A useful design matrix illustrates ‘HOW’ to achieve sustainable integrated design and the way it is applied in most areas of building design.

4.1 Design with a Matrix

67

‘WHAT’ and ‘WHO’ “Conscious architecture, as distinguished from vernacular building, should be able to reason out the unique solutions to specific problems” (Banham 1969). Reyner Banham in his book, The Architecture of the Well-tempered Environment emphasised how, from the industrial revolution onwards, multiple parameters became part of the new design process, contributing to the environmental quality of buildings. “It would have been apparent long ago that the art and business of creating buildings is not divisible into two intellectually separated parts—structures, on the one hand, and on the other mechanical services” (Banham 1969). At a time when experimentation with new materials, new buildings, and new ways of living and working are the stimuli behind research and design, performancebased building design and human-centred design help to satisfy not only the needs of users, but also those of the actors involved in the building process, from designers to citizens. The development of new building techniques and new functional paradigms needs to be coupled with an analytical approach to design, one which is more than just a simple aesthetic-compositional exercise and involves consolidated building techniques that are the result of sedimented and experimented experiences (Arbizzani 2015). This implies that buildings are considered as a system in which components are no longer separate elements characterised by a technical and functional performance, but together satisfy a requirement. “If he [the architect of the future] will build up a closely co-operating team together with the engineer, the scientist and the builder, then design, construction and economy may again become an entity—a fusion of art, science and business” (Gropius 1962). In the fifties Walter Gropius stated that, unlike industry, in architecture the prototype and final product are one and the same. It is therefore possible to define a building as the unique, complex result of a “building chain”. The chain creates a systemic approach by merging both the product chain concept typical of industry and the requirements of the actors involved in the life cycle of the building. In the Life Cycle Design approach the performance characteristics of each individual part interact and at times interfere, creating a multidisciplinary integrated design process. But how is it possible to control this complex design process? It cannot be linear in terms of analysis and timeline because the dynamic interaction between the actors, stages and knowledge has to be governed. A tool is required to link needs, requirements, and performance; it has to be scalable depending on the evolution of the project; it must consider not only the design stage, but also the construction of the object, its use and long-term management, as well as possible scenarios at the end of its service life cycle. This can be achieved by using one or two matrices to not only represent the requirements of the actors involved and the performance characteristics of parts of the building, but also visualise and solve the interactions and interferences that may arise, including with the support of different disciplinary fields. In the early design stages, client and designer begin by establishing a ‘picture’ of their needs and constraints; then, based on this picture they create the first matrix with the ‘rules’ of the design. Afterwards they involve both the ‘specialists’ (e.g., structural engineers, MEP engineers, geo technicians, historians, urban planner, final users, and

68

4 Defining an Innovative Design Method …

Fig. 4.2 Conceptual map of performance-based building design

other specialists) and ‘non-experts’ (e.g., interested citizens, local authorities, etc.). The number and role of these specialists and non-experts is established based on the type of project (use, link with the site, etc.); together they elaborate the needs and requirements of the feasibility study. At this point, the team coordinator can establish the design guidelines and produce the final design which will be shared, verified and re-verified with the other actors involved. This will continue throughout all the design stages until the performance indicators have been calculated (Fig. 4.2). ‘HOW’ How do the actors exchange and share information? How can a design be based more on human needs than on the building? How can the future users of the building be empowered and educated? The task of the design team is not only to develop and send specialists a huge amount of complex data in a simple, correct manner, but also involve ‘non-experts’, including the users directly affected by the service life cycle of the building. The goal of Life Cycle Design is to guide and empower the actors in a “living” process that starts with the design and continues with the use of the asset based on the envisaged “life cycle of the building”. Many actors will spark this process. First the clients, then the designers, and finally the users. A matrix has to be developed to establish strategies which then need to be shared and shaped together with the stakeholders

4.1 Design with a Matrix

69

Fig. 4.3 Example of the two-dimensional matrix. The life cycle stages (time) are located along axis y, while the requirements of each stakeholder are along axis x. The design strategies are created in the area where the two axes overlap

involved in the design process. The design matrix must be developed from the early design stages to the end of the process. A final version of this matrix must include the performance indicators taken into consideration and the envisaged life cycle of the building. Both factors vary according to the design. This tool cannot and must not replace the creative act leading to a unique, complex design. An architectural design is not a mathematical formula with operations following on in a univocal sequence. However, the analytical nature of the matrix is a crucial element that has to go hand in hand with the creative act in order to enhance the interdisciplinary exchange of knowledge, the integrated elaboration of technical parameters, respect for the timeline, and communication of the design. The matrix is initially two-dimensional and includes the design strategies aimed at satisfying the functional, regulatory, environmental, social and economic needs behind the design procedure. Each life cycle stage (in Fig. 4.3, axis y  time) is linked to certain needs and requirements of the actors involved in the design process (in Fig. 4.3, axis x  needs, requirements). The first elaboration of the matrix (the concept design matrix) is based on qualitative assumptions and represents the sharing of the performance-based framework between the client and the architect; the second elaboration (the developed design matrix) is instead shared with all the users involved. All the stakeholders can elaborate their own design matrix that can be shared with all the team. The matrix (the technical design matrix) becomes more complex as the design evolves: needs and requirements are used to quantitatively define the quality/sustainability performance indicators for each life cycle stage (in Fig. 4.4, axis y  time; axis x  indicators). The initial strategies are gradually represented by the performance levels to be achieved through quantity (in Fig. 4.4 axis z  quantity). The matrix then becomes three-dimensional (indicator of quality/sustainability, quantity, time).

70

4 Defining an Innovative Design Method …

Fig. 4.4 Example of the three-dimensional matrix. The life cycle stages are along axis y, the performance indicators along axis x, the value associated with each indicator (quantity) along axis z

In this kind of structured design process it is easier to maintain coherent links between the choices made. It is also possible to achieve optimisation during both the design stage and use of the building. Each performance indicator can be analysed as disaggregated data either to perform a more in-depth analysis of specific issues, or to compare different options (e.g., selecting the most efficient air conditioning system or the most suitable building system compared to the established performance-based framework, etc.).1 Despite the fact that the value scale of the different indicators varies according to the unit of measure, the overall matrix summarises the design parameters: which, how many, and during which stage certain aspects have been considered and will produce effects. The next chapter outlines the calculation method of the evaluation model based on the eCO synthesis factor. The factor combines the environmental and energy effects of the building system with the time and cost factor. The design matrix indicators are the input data of the Life Cycle Design Model illustrated in the following chapters.

4.2 The Life Cycle Design Model (LCDM) The trade-off between validity and applicability of a model is a big challenge (Finkbeiner et al. 2010). It begs the question: how can the inherent complexity of this approach be transferred to a non-expert audience and real world decision-makers? This chapter illustrates a second level of assessment, which consists in the development of a user-friendly tool for designers and other actors involved in the building 1 The use of integrated design tools, for example BIM, Building Information Modelling, facilitates the application of the matrix. In fact, by using BIM the matrix can represent a concise graphic image of the calculations performed.

4.2 The Life Cycle Design Model (LCDM)

71

Fig. 4.5 Example of the three-dimensional matrix. A quantitative value, divided according to the life cycle stages, is associated with each indicator. The quantities are expressed by different units of measure. The dotted line shows the indicators analysed with the LCDM (data input)

process so that they can evaluate the most sustainable design option: the Life Cycle Design Model (LCDM). Through points and weights it expresses an economicenvironmental efficiency value (eCO), one which is more subjective and controversial then the design matrix based on separate indicators. However, it can be a useful tool in the design process to share information with specialists and non-experts (clients and communities). The LCDM can take into account some or all of the parameters in the Design Matrix, for example energy balance, environmental impact, comfort, costs, aesthetic value, historical value, integration in the urban, social and cultural context, and other indicators (see the case studies illustrated in Chap. 5). Figure 4.5 shows the threedimensional matrix with the indicators and impact divided according to the life cycle stages (in Fig. 4.6 axis y  time; axis x  indicator of quality/sustainability; axis z  quantity). This study considers three indicators: Life Cycle Assessment, Life Cycle Costing, and energy performance. Life Cycle Assessment (LCA) quantifies the environmental and energy impact of the building materials and construction components, structures, systems, etc.2 It includes the impact of the following stages: construction, including the extraction of raw materials, production of the materials, and their installation; maintenance of the 2 See

Chap. 3 for more details on LCA.

72

4 Defining an Innovative Design Method …

Fig. 4.6 The diagram represents the life cycle approach; starting with design strategies it shows more in-depth data quality

building, including replacement of elements and materials; and end of life either of building elements or the building itself. Energy performance represents the envisaged quantity of energy needed to make the building functional and can include winter heating and summer cooling, artificial lighting, mechanical ventilation, and the production of sanitary hot water. In other words, energy performance expresses, as Operational Energy, the energy impact of the building and systems during use, and as Embodied Energy and Recurring EE, the energy impact of the building materials during construction and use. Life Cycle Costing (LCC) must ensure that all the costs arising from the replacement of a product or system during the life cycle be integrated into the decisionmaking process from the very beginning so that logical choices may be made to avoid environmental damage and negatively affect society (Molinari 2002; Hunkeler et al. 2008). This premise leads to the conclusion that the cost analysis must be applied to the entire life cycle in order to represent the concept of sustainability.3 This involves establishing how to coherently combine costs and energyenvironmental aspects (including the effects caused by society). Critical aspects and obstacles related to this issue will be studied in the first part of the book; the research goal illustrated in this chapter aims at solving the issues associated with the integration of LCC and LCA models in the field of construction.4 Furthermore, before describing the method it is important to specify that the latter can be applied from the very early design stages when design strategies are defined. The accuracy of the model depends strictly on the level of definition of the design (Fig. 4.6). If, for example, it is applied during the early design stages it will be impossible to use the primary data referring to the materials, products, and their quantities. However it will be possible to use qualitative or parametric indicators based on the data available for that specific design stage. For example, if a wooden structure is envisaged, it can be assessed and compared with other technologies using 3 See

Chap. 3 for more details on LCC.

4 Always bearing in mind the suggestions expressed in ISO 14040: “LCA typically does not address

the economic or social aspects of a product, but the life cycle approach and methodologies described in this International Standard may be applied to these other aspects”.

4.2 The Life Cycle Design Model (LCDM)

73

qualitative indicators based on a range of values regarding, for example, the weight, availability and source of the material, and maximum size of the elements. The more in-depth the design, the more detailed the data regarding the materials, components and building system and, therefore, the more accurate the LCDM result.5

4.2.1 Basic Principles of the Life Cycle Design Model To produce an “easy-to-communicate” result, the Life Cycle Model associates the costs of a product (building or component) with its environmental impact and energy performance based on the data and information in the Design Matrix. The goal is to develop a tool which, based on complex, multifaceted data, can help designers make design choices (e.g., select the building system, materials, assembly types, etc.), allow private clients to compare different environmental efficiency technologies (environmentally-friendly materials, greater durability of the elements, etc.) and, finally, be useful to public authorities during assessment (e.g., tenders). Both expert and non-expert users must be able to use the tool. This decision-making approach aims at defining more intelligent, informed design choices, the effects of which will, albeit indirectly, affect society. The specific goals of the model are summarised as follows: 1. Quantify and compare the environmental, energy and economic performance or so-called global performance6 of alternative solutions regarding components and the building. 2. Provide a common calculation method that can be inserted in multidisciplinary tools (e.g., BIM). 3. Ensure flexible use and scalability of the tool according to the goal of each project and design scale. 4. Provide private investors or public authorities with a control tool during the design stage. 5. Allow all the actors (owners, suppliers, designers, etc.) to make decisions based on a quality/price ratio (including the “quality” of environmental aspects). 6. Make it possible to compare alternative technological systems vis-à-vis the entire life cycle highlighting the most relevant life cycle stages. 7. Minimise the risk of uncertain results by using a sensitivity analysis. Figure 4.7 illustrates the hypothetical system made up of an ensemble of input and output data. The input data constitutes all the energy, environmental and economic performances throughout the life cycle summarised in the Design Matrix. These 5 Chapter

5 illustrates several examples with different levels of design detail. E.g. case study 5.1.1 (VM House) is based on qualitative parameters, while case study 5.2.1 (the Chavonne warehouse) focuses on data in the final design. 6 “Global performance” is based on the merger of several quality indicators and envisages the direct involvement of actors from other disciplines throughout the design process.

74

4 Defining an Innovative Design Method …

Fig. 4.7 Functional diagram of the Life Cycle Design Model (LCDM)

assessments, carried out using computer tools,7 are expressed according to the unit of measure of the performance scales (e.g., e, m2 , W/m2 K, etc.). The output data represents the global performance values of each impact indicator normalised to a single functional unit (e.g., gross floor area, area of a technical element, etc.). The model enables assessment of the environmental and economic performance of a building or parts of it during its entire life cycle based on the limits established by the user during the phase known as “defining the scope of the analysis and reference framework”. A sine qua non condition of the Life Cycle Design approach is to consider all the life cycle stages, because selecting a building type based only on a single life cycle stage would mask possible environmental damages, either similar or greater, that may be generated during other stages. As mentioned earlier, the results of the environmental, energy and economic assessments constitute the INPUT data of the model. The three performances are later elaborated using a calculation algorithm which will provide the overall performance that can be used to compare technological systems. The outcome obtained for each alternative design is represented by the economic-environmental factor, in other

7 See

Chap. 1, paragraph “Integrated design tools”.

4.2 The Life Cycle Design Model (LCDM)

75

words the concise indicator called eCO, corresponding to values between 0 and 1 compared to the benchmark. Transparent results are ensured by the fact that the results of the economic, environmental, and energy analyses can be used either as aggregated or disaggregated data using a system of weights. Selecting weights varies according to the priority of each design. Double representation of the results of the proposed method intends to guarantee not only greater validity of the system,8 but also maximum applicability, thereby ensuing that non-experts can understand the results. Several studies carried out by the scientific community (König et al. 2010; Mateus and Bragança 2011; El Khouli et al. 2015; Pombo et al. 2016a, b) show that simplification is required in order to disseminate this approach. However, ‘simplification’ does not mean a lack of scientific validity, because simplification is based on knowledge and the elaboration of very complex data. Proof comes in the form of further in-depth study during interpretation of the results; the suitability of the tool was verified, focusing in particular on the criticalities identified during a study of the state of the art regarding simplification of assessment tools. In particular, three verification methods were adopted9 after a comparison of the tools cited in literature: 1. Graphic representation (Method A) The energy-environmental impact (axis of the abscissa) and costs (axis of the ordinates) are compared in a bubble graph.10 2. Curve method of the marginal CO2 abatement cost (Method B) The marginal abatement of carbon dioxide for alternative building solutions is calculated compared to the benchmark. The result is based on the hypothesis that emissions trading exists in the building market.11 3. Method to monetise the environmental impact (Method C) The environmental impact is internalised by monetising the environmental performance which will then be added to the economic performance.12 Developing the Life Cycle Design Model was inspired by the study of the interdisciplinary framework described in the first part of the book. The following list of the international experiences focusing on integration between LCA and LCC includes, amongst others: • ISO 15686-5 “Building and constructed assets—Service life panning—Life cycle costing” (published in June 2008, last update 2017). • “Environmental LCC” in Hunkeler D., Lichtenvort K., Rebitzer G., Environmental Life Cycle Costing, SETAC Books, CRC Press, NY (November 2008). 8 As

recommended by UNI EN ISO 14044. Sect. 4.4. 10 “Method A” was inspired by both the “Ashby bubble plot theory” (CES—Granta Design, UK) and the “Relative life cycle portfolio” (Environmental LCC—Swarr T., Hunkeler D. “Life Cycle Costing in Life cycle management”, 2008). 11 “Method B” is developed using the economic principle of the emissions tax (cf. Sect. 3.5.2) and is based on the Emission Trading market mechanism (a system for greenhouse emissions allowance trading—European Directive 2003/87/CE). 12 “Method C” is based on monetisation factors defined by the Swedish impact evaluation method EPS 2000 and on the “ExternE” project sponsored by the European Union in 1999. 9 See

76

4 Defining an Innovative Design Method …

• ISO 14040 “Environmental management—Life cycle assessment—Principles and framework” (published in 1997, last update 2006 confirmed in 2018) and ISO 14044 “Environmental management—Life cycle assessment—Requirements and guidelines” (published in 2006, confirmed in 2018) • Building for Environmental and Economic Sustainability (BEES)—National Institute of Standards and Technology (NIST) • Davis Langdon Management Consulting, “Life cycle costing (LCC) as a contribution to sustainable construction: a common methodology”. May 2007 • EMSD and Arup, “Life Cycle Assessment (LCA) and Life Cycle Cost (LCC), Tool for Commercial Building Developments in Hong Kong”. June 2006 • Task Group 4: “Life Cycle Costs in Construction. Final Report”. July 2003

4.2.2 General Framework The structure of the model was inspired by the one defined in ISO 14040 for LCA analysis. The following are the steps required (Fig. 4.8): STEP 1 STEP 2 STEP 3 STEP 4 STEP 5 STEP 6

Definition of the goal and scope Inventory analysis (LCI) Impact assessment (LCIA) Disaggregated results of the comparison of design options (tot LC) Individual results of the comparison of design options (eCO) Interpretation of the results.

The structure of the Life Cycle Design Model is illustrated here as an example of how the model is used to assess the economic-environmental impact of opaque envelope elements. The example is taken from the author’s research submitted as

Fig. 4.8 Framework of the life cycle design model (LCDM)

4.2 The Life Cycle Design Model (LCDM)

77

part of her Doctorate in Technological Innovation for the Built Environment, XXI cycle, presented in 2009.13 Despite the fact that the functional unity, the system boundary, or the type of impact, can vary depending on the objectives of the study and design, this chapter will use a specific example to illustrate the general methodological process. Nevertheless, since the objective of the research is to demonstrate the validity of the method using tangible, diversified applications, a complete interpretation of this design approach is provided in the last chapter illustrating case studies, their objectives, functional units, and the boundaries of the various systems. STEP 1: Definition of the goal and scope As specified by several of the international studies taken into consideration (UNEP 2011), in order to compare LCA and LCC coherent goals and scope have to be identified by defining the scope of the analysis, system boundary, and functional unit. An envelope was used as an example. The aim of the model is to perform an economic-environmental assessment of alternative designs; the assessment is based on the LCC method to estimate costs and on the LCA method to evaluate the environmental and energy impact. Several building technologies14 were compared with local standard technology (thermal resistance being equal) in order to determine the best solution in terms of environmental quality and global cost during the life cycle. Figure 4.9 illustrates the stratigraphies that were analysed. The standard solution represents business as usual (BAU),15 i.e., the benchmark for the comparison with other opaque building envelope solutions. Appendix 2 contains the data of the case study.

4.2.2.1

System Boundary

Defining the scope of the analysis is closely linked to the identification of the system boundary. In this case “system” means all the processes activated throughout the life cycle of the building. There are two scales regarding the boundary of the system: 1. The “building” scale. Figure 4.10 illustrates how the building system can be divided into: construction system, load-bearing structures, and systems (MEP—Mechanical, electrical, and plumbing). The boundary of the described model is shown by a blue dotted line 13 The

data contained in the case study refer to 2008. Thiébat (2009), Architettura e sostenibilità: sviluppo di un modello economico-ambientale. PhD dissertation, XXI cycle, Tutor: M. Grosso, Politecnico di Torino. 14 The alternative building technologies used in the case studies were chosen based on their characteristics of eco-compatibility and easy assembly/disassembly. 15 The reference building technology of the standard case study will be indicated with a term borrowed from economics: “Business as Usual”. It will represent the typical building method used in the Piedmont-Valle d’Aosta region (Italy).

78

4 Defining an Innovative Design Method …

Fig. 4.9 Example of alternatives to the opaque envelope compared with the LCDM model

4.2 The Life Cycle Design Model (LCDM)

79

Fig. 4.10 System boundary for the building (example)

(technical element 01  opaque envelope). This boundary can also be broadened to include other elements or the whole building, as per the examples illustrated in the next chapter. 2. The “Process” scale This indicates which stages and which processes have been considered. Several processes included in the study of the example are provided below: • extraction of the raw material/transportation to the production site; • transformation of the raw materials and production of the materials for the envelope of the building; • supply and installation of the envelope materials for the construction stage; • cyclical maintenance of the envelope materials during use of the building; • replacement of the materials or components depending on their functional life cycle (removal and installation of new materials/components); • possible reduction of the energy performance of the building due to loss of performance of the construction system; • recycling/disposal of the materials in a landfill. The example does not include transportation from the production company to the worksite and the energy needed to install the envelope.

4.2.2.2

Functional Unit and Flowchart

As mentioned earlier, it is important to identify a functional unit common to both the LCA and LCC. The parameter adopted in the example is one square meter of technical element 01 (1 m2 of a vertical opaque envelope). A flowchart has to be created to establish the system boundary and identify the characteristics of the model; not only does it have to show the processes that can be

80

4 Defining an Innovative Design Method …

Fig. 4.11 Reference flows used to create the system

used to collect the inventory data, but also the amount of material (e.g., in kilograms), energy (e.g., kWh), and relevant costs (e.g., in Euro). Figure 4.11 shows the standard diagram for the creation of the reference system of the model: all the material and energy, whether entering or exiting, corresponds to an economic flow. In order to be able to compare the alternatives and include energy performance in the stage of use, it is particularly important to normalise the functional unit to the technical performance (in the example, 1 m2 of the surface of the construction element with the same R-value). In this case, to select the best solution (thermal transmittance being equal) normalisation was applied by dividing the impact indicators by the thermal resistance value (R-value) of each technical element and then multiplying them by a constant R-value of 2.5 m2 K/W. This procedure makes it possible to compare the impact of the different building technologies while maintaining a single unit of measurement (energy performance being equal). As concerns the duration of the life cycle, the proposed assessment model considers a period of analysis as close as possible to the service life envisaged for traditional local building products, in other words 70 years (75 years is the indicative duration of an outdoor brick wall, insulated and with an installed vapour barrier).16 As mentioned earlier, the presumed duration used for the period of analysis should not be confused with the life cycle duration of the material, the technical element, or the building, because they may not coincide. In fact, literature informs that the period of analysis must be between 40 and 80 years if the analysis is to be effective (Molinari 2002; BEES 2006; DGNB; Dodd et al. 2017). Shorter periods will not be significant from the point of view of the depreciation of the initial costs, while longer periods will cause the initial data and, hence, the results obtained, to be unreliable, especially as regards three factors: technological obsolescence; excessive uncertainty regard-

16 Cfr. Molinari (2002). Appendix 2 illustrated the data regarding the life cycle duration of building

subsystems specified in: Dell’isola, Kirk, Life Cycle Cost data, McGraw-Hill, NY, 1983, based on the American ASTM, Building Maintenance, Repair and Replacement Database for Life Cycle Cost Analysis (dated 1995).

4.2 The Life Cycle Design Model (LCDM)

81

ing the data; “the longer the payment, the lesser the importance given to the costs” (BEES 2006). Using the division of phases indicated in Chap. 3, the Life Cycle Design Model was developed based on three life cycle stages: 1. product stage: • acquiring raw materials • production of building materials and components 2. construction stage: • construction of the artefact 3. use stage: • use of the artefact • ordinary maintenance • extraordinary maintenance and replacement 4. end of life stage: • decommissioning (disposal and/or recycling)

4.2.3 The Environment and Energy Factor The environmental assessment is based on the “from cradle to grave” approach or, in some cases, “from cradle to cradle”. This means that the environmental impact should be considered in every end of life stage until decommissioning of the object of the study: the building or building component. Energy assessment is divided into the primary energy used to build/demolish the building and produce, assemble/disassemble and replace its components, and the primary energy needed during the use stage. Figure 4.12 refers to the example of the building envelope. The life cycle begins with acquisition of the raw materials, transportation, and production of the materials. It continues with the use stage which includes maintenance and replacement of individual components of the building element according to the guaranteed duration of the material, and terminates with the end of life stage, disposal or reuse. The construction stage has not been included in the environmental assessment because it depends on the type of worksite, as specified by the definition of the system boundary. Since the model is based on the functional unit of 1 m2 of the technical element (thermal resistance being equal), it would in fact be contradictory and reductive to theorise a fixed scenario by defining the amount of material, location of the worksite, and use of the building. Nevertheless, it is possible to include these

82

4 Defining an Innovative Design Method …

Fig. 4.12 Environmental impact of the life cycle stages (envelope example)

items in the system boundary in order to perform a more detailed comparison based on available design data. STEP 2:17 Inventory analysis (LCI) This step includes data collection and the calculations quantifying the input and output flows of a product system in terms of the energy and materials taken from the environment, and the emissions and waste re-entering the environment. Each phase of the LCA study contains “primary” and “secondary” data. In the envelope example primary data was collected directly on the worksite and refers to the metric calculation and data either provided by the producers or taken from the LCA and EPD studies performed by the DAD18 research team (Giordano 2010). Secondary data was collected from European databases, especially EcoInvent. An inventory analysis was performed for each stage and each alternative after selecting the construction types to be compared. Table 4.1 indicates, for example, the data of the product/construction stage of a technical element as well as the amount of materials in each strata. This procedure was applied to the other life cycle stages and to every stratigraphy identified in the research project (cfr. Appendix 2 to this chapter). STEP 3:19 Environmental impact assessment (LCIA) This step involves processing the results of the inventory. The objective is to assess potential environmental impacts and therefore underscore not only the extent of the environmental changes caused after emissions or waste re-enter the environment, but also the resources used during production. Sima Pro 7.1 software was used to assess the impact of the envelope and construction of the LCA model. The following methods were used for impact factors characterisation20 : • single indicator method to calculate the Global Warming Potential GWP: EPD 2007 and IPCC 2007. 17 See

Fig. 4.8: green line step 2. of Architecture and Design, Politecnico di Torino. 19 See Fig. 4.8: green line step 3. 20 See Chap. 3 for the impact factors. 18 Department

4.2 The Life Cycle Design Model (LCDM)

83

Table 4.1 Example of an inventory table with the data collected for the BAU stratigraphy Building materials

Layer thickness (m)

Weight density (kg/m3 )

Weight (kg/m2 )

Thermal resistance (m2 K/W)

Quantity (for a component surface of 1 m2 ) (kg)

BAU

Cement plaster

0.005

1800

9

Cement render

0.010

1800

18

Clay-brick

0.120

1500

180

Clay-brick

0.080

775

62

EPS panel

0.100

15

1.5

1.50

Gypsum plaster

0.015

1400

21

21.00

Thin-coat gypsum plaster







Wall paint

2 layers



0.16

Total

0.33

7290

291.66

27.00

242.00

0.16 2.82

291.66

• single indicator method to calculate the Gross Energy Requirement GER: CED v. 1.01. The values obtained from the environmental impact assessment constitute the environmental INPUT data of the LCDM model (Fig. 4.7). STEP 4:21 Disaggregated results in order to compare design options (tot LCA). To compare the design alternatives, the impact factors with their relative units of measure were normalised to the functional unit FU of the study (Eq. 4.1). This makes it possible to calculate the environmental assessment for each impact k (tot LCAk ). In the example the functional unit FU corresponds to 1 m2 of the envelope with a Thermal Resistance equal to 2.5 m2 K/W. tot LC Ak 

qk F.U.

[um k /F.U.]

(4.1)

where: tot LCAk qk F.U. k umk 21 See

Environmental assessment relating either to the environmental impact k Flow of the environmental impact k Functional Unit Type of impact (e.g., GWP; GER) Unit of measure of the k impact (e.g., kg of GWP; MJ of GER).

Fig. 4.8: green line step 4.

84

4 Defining an Innovative Design Method …

Table 4.2 The construction stage of the benchmark envelope solution (BAU)—inventory data and impact assessment associated with the Global Warming Potential (GWP) and Gross Energy Requirement (GER) (data processing by SimaPro 7.1). The values refer to 1 m2 of the envelope with a thermal resistance equal to 2.5 m2 K/W Building materials

BAU

Process 1 | Construction stage Environmental database

Waste type

Ecoinvent

Ceramics

27.00

Ecoinvent

Ceramics

242.00

EPS panel

Ecoinvent

Plastics (type 6)

Gypsum plaster

Ecoinvent

Ceramics

21.00

Ecoinvent

Ceramics

0.16

Cement plaster

Quantity (for a com poneit surface of 1 m2 ) (kg)

Total GER process 1 (MJ/m2 )

Total GWP process 1 (Kg CO2eq /m2 )

642.51

166.43

Cement render Clay-brick Clay-brick 1.50

Thin-coat gypsum plaster Wall paint Total

291.66

The results of the environmental assessments are inserted in the impact technical sheets. Table 4.2 shows the source of the inventory data, the type of waste, and the quantities of the BAU solution. It also provides the results of the impact analysis related to the construction stage. The same procedure was applied to other alternative envelopes for each life cycle stage (cfr. Appendix 2 in this chapter).

4.2.4 The Economic Factor Measuring the economic factor may seem easier compared to calculating environmental performance. Data is relatively available and the calculation method was normalised by the ISO. Nevertheless, as illustrated in earlier chapters, life cycle costs are very difficult to assess and vary significantly depending on the inflation and rebate rate adopted with regard to the envisaged duration of the material since this influences the number of replacements and/or maintenance cycles during the service

4.2 The Life Cycle Design Model (LCDM)

85

life of the building and, as a result, the costs of maintenance. In other words, the role of “time” is certainly crucial in any economic assessment since it can lead to uncertainty and random results. As mentioned in chapter one, it is important, first and foremost, to distinguish between the period of analysis of the economic assessment and that of the environmental assessment. The goal is to analyse the role from different viewpoints: for a private investor the period of analysis corresponds to the length of the investment, while a public company would consider that the period runs, for example, for the entire service life of the building (Hunkeler et al. 2008). However, these hypotheses are based on the need to maintain a single period of analysis covering all the life cycle stages. After expressing these considerations, and as per the environmental assessment, the economic factor focused on: the categories of costs; the assessment methods to be used during the inventory phase and during characterisation. STEP 2:22 Inventory analysis and classification of costs According to ISO 15686-5 the LCC method combines all the important costs of every building element, costs which will have to be borne during the period of analysis. The economic factor is obtained by summing the costs calculated with the LCC method, based on the following equation: LCC  Cs + Ci + Cm + Cr + Co + Coel where: Cs Ci Cm Cr Co C eol

Supply costs Installation costs Maintenance costs Replacement costs Operational costs End of life costs.

Figure 4.13 shows the economic impact axis with the costs of each life cycle stage for the building envelope. Supply costs (Cs ) and installation costs (Ci ) were considered for the construction stage; these costs were based either on the market price of the materials that make up the stratigraphy23 or on suppliers’ estimates. Maintenance costs (Cm ) and/or replacement costs (Cr ) of one or more strata of the technical element during the use stage were extrapolated from price lists or handbooks24 and discounted future costs to 2008, depending on the year when the element was replaced. If it was impossible to replace a strata, then the “loss of performance” cost (operating costs Co ) was calculated taking into consideration the kWh of thermal energy consumed due to the increase in heating requirements (e.g., a hollow wall 22 See

Fig. 4.8: blue line step 2. price list of the Regione Piemonte (Italy), Prezzario di riferimento per opere e lavori pubblici nella Regione Piemonte (ed. 2008). 24 C. Molinari, op. cit. 23 Source:

86

4 Defining an Innovative Design Method …

Fig. 4.13 Life cycle stages vis-à-vis the economic impact (example)

with insulation between the bricks25 ) and based on a hypothetical