Creativity, Problem Solving, and Aesthetics in Engineering: Today's Engineers Turning Dreams into Reality 3030382567, 9783030382568

This book illuminates what engineering is and how it relates to other disciplines such as art, architecture, law, econom

119 6 8MB

English Pages 242 [234] Year 2020

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Creativity, Problem Solving, and Aesthetics in Engineering: Today's Engineers Turning Dreams into Reality
 3030382567, 9783030382568

Table of contents :
Preface
Reference
Praise for Creativity, Problem Solving, and Aesthetics in Engineering
Contents
1 Dreaming
Ben’s Story
What Is Engineering?
A Clarification
Why Is Engineering Important?
Who Are Our Engineers?
All Is Not Well on Planet Earth
2 Making
Craft Making
Art Making
Functional Art Making
Invention
Engineering Making with Science
Engineering and Technology
3 Grounding
Five Principles
The Clockwork Universe—A Discredited Myth?
A Learning Odyssey
Problem Solving with Practical Wisdom
4 Dwelling
Ronan Point and Grenfell Towers
Shelter
Theory and Practice
Fragmentation
Architectural and Structural Form
How to Read a Structure
Wind and Vibrations
Progressive Collapse and Terrorism
Structural Safety
5 Moving
Edith’s Story
Moving
Early Land Transport
Water Transport
Land Transport
Engines
Fluids and Solids
Heat Engines
Turbines
Jet Engines
Moving On
6 Communicating
Patterns
The Syntax of Jacquard’s Loom
Signalling
Michael Faraday
James Clerk Maxwell
Transmitting the Human Voice
Computing
Semantics and Pragmatics
Perspectives
Passive and Active Information
Big Data Becomes Active
Artificial Intelligence
Women in AI
Getting Smaller
7 Fighting
Literary Engineering
War
Weapons
Trebuchet—Early Ingenuity
Gunpowder
Samuel Colt
Baron William Armstrong
Radar
Nuclear Weapons
Rockets and Missiles
Targets
Defence
Sustainability
8 Well-Being
Robert Langer’s Story
Well-Being
Agricultural Making
Drinking Water
Waste
Joseph Bazalgette
Pacemaking
What Next?
9 Flourishing
West Gate Bridge
Will This Never Happen Again?
Revisiting Patterns
Practical Wisdom and Rigour
Grand Challenges
Engineers at the Heart of Society
Meeting the Challenges
Fragmentation
Learning Together Requires Leadership
Creativity, Problem Solving and Aesthetics—Turning Dreams into Reality
Engineering Is a People Profession
Glossary
Index

Citation preview

David Blockley

Creativity, Problem Solving, and Aesthetics in Engineering Today’s Engineers Turning Dreams into Reality

Creativity, Problem Solving, and Aesthetics in Engineering

David Blockley

Creativity, Problem Solving, and Aesthetics in Engineering Today’s Engineers Turning Dreams into Reality

123

David Blockley University of Bristol Bristol, UK

ISBN 978-3-030-38256-8 ISBN 978-3-030-38257-5 https://doi.org/10.1007/978-3-030-38257-5

(eBook)

© Springer Nature Switzerland AG 2020 This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. The publisher, the authors and the editors are safe to assume that the advice and information in this book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty, expressed or implied, with respect to the material contained herein or for any errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. This Springer imprint is published by the registered company Springer Nature Switzerland AG The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland

One of the great revelations of the space age has been the perspective it has given humanity on ourselves. When we see the earth from space, we see ourselves as a whole. We see the unity and not the divisions. It is such a simple image with a compelling message— one planet—one human race. We are here together, and we need to live together with tolerance and respect…. Our only boundaries are the way we see ourselves. The only borders, the way we see each other. Let us fight for every woman and every man to have the opportunity to live healthy, secure lives, full of opportunity and love………. We are all time travellers, journeying together into the future. But let us work together to make that future a place we want to visit. Stephen Hawking World Economic Forum 2015 Transcribed from https://www.weforum.org/agenda/2015/09/ stephen-hawking-we-are-all-time-travellers/ The test of our progress is not whether we add more to the abundance of those who have much it is whether we provide enough for those who have little. Franklin D. Roosevelt USA Presidential Inaugural Address, 20 January 1937.

Preface

Cave dwellers had a life expectancy at birth of around 19 years. Today a baby is expected to live to well over 70 depending on the country of birth (World Health Organisation 2015). This book is about the engineering of that transformation—the way we humans have dreamed and ingeniously turned new ideas, and new ways of doing things, into reality. Engineering is not only something-to-do-with-engines. We engineer when we are inventive, resourceful and aesthetic. The word engineer stems from the Latin ingeniator meaning someone who is ingenious. The modern French word is ingénieur. To engineer is to solve problems as only we humans can, mindful of our delight in our creativity. In that sense, we are all engineers. Before the Renaissance, the same person might be artist, artisan, architect, craftsman, mason or engineer depending on the job he was doing—the distinctions we make nowadays between aesthetics and function were small. The things we engineered like wagons, ploughs, cannons and even cathedrals were relatively easy to grasp. We still look in awe at the sheer size of the Egyptian pyramids and the beauty of the towering vaults of a gothic cathedral and we know individual masons, craftsmen and others developed particular ingenious skills. But we can appreciate something as to how they did it without us having to have any scientific education —after all, they worked intuitively long before the science we have now. Today circumstances are quite different. Engineers have high levels of education and training and specialised knowledge and skills to make the buildings, roads and bridges, cars and trains, ships and aeroplanes, TVs, radios, computers and smartphones that we rely on. They make the ‘systems’ of things such as airports and urban infrastructure. Their professional duty of care for the safety and well-being of others requires them to examine the scope and dependability of all kinds of information, including science, as they make useful and often challenging and risky things. Through the ages, the relentless, often dramatic, erratic and ever accelerating, unfolding of change has been impressive. But now some of the things they make are so clever and complicated that most of us have little or no idea how they work so we hardly give them a thought until they fail. How many of us know what

vii

viii

Preface

is going on inside the ‘black box’ of a computer or the jet engine of an aeroplane? They are totally opaque and yet we rely on them—we have little choice. In the twenty-first century, we have become collectively much more aware of some big concerns. All is not well on planet Earth. We have been so successful that we have made ever-increasingly powerful ways of killing each other. We have caused pollution and we are changing our climate. We have created social media that helps us to talk and connect with others but also allow the unscrupulous to exploit the vulnerable. As we contemplate an uncertain and, in many ways, disquieting future, we owe a duty to future generations to take stock of what has been and what might be forthcoming. In Greek mythology, the god Zeus gave Pandora a box with strict instructions not to open it. When curiosity got the better of her, all the evils and miseries of the world flew out. Richard Sennett says in his book The Craftsman (2008), ‘The contents of Pandora’s Box can indeed be made less fearsome; we can achieve a more humane material life, if only we better understand the making of things’. Richard Sennett continues ‘Learning from things requires us to care about the qualities of cloth or the right way to poach fish’. Buckminster Fuller expresses it as an aesthetic ‘where the stresses and strains are at ease’. Steve Jobs is quoted as saying ‘aesthetic is the quality that has to be carried all the way through’. In this book, I aim to do three things. First to expand on these affirmations of aesthetics in practice by showing how engineers have made and still strive to make the quality of our lives better. Second to identify and explore some of the unintended consequences of the past and the ‘grand’ challenges ahead. Third I will suggest some ‘grounding’ principles that may help us to guide or steer our way through a risky future. The book has been written primarily for non-technical readers who may or may not be makers but who do not normally think of themselves in any way as engineers. Technically trained engineers wishing to look beyond the detail of their primary discipline may find some congruity. Likewise, there may be benefit for business leaders, lawyers, economists, politicians and all others who must make schemes and stratagems for making practical decisions. As I have already hinted the perception of professional engineering by the wider world seems to be scant and patchy according to the little research that has been done. An article by Nadya Anscombe in the New Scientist in 2005 is perhaps typical: ‘It can be hard to convince outsiders that engineering is a worthwhile profession. Perhaps that’s because the name ‘engineer’ has come to be attached with men (rarely women) in overalls, fixing phones or mending bits of machinery……It’s a familiar story for engineers: you are at a dinner party or family gathering and someone asks you what you do. When you tell them, their eyes glaze over and they quickly change the subject’. Many people associate engineering with engines, mechanisms and machines. Synonyms for the word ‘mechanical’ include the words cold, emotionless, impersonal, routine, monotonous, unfeeling, unthinking and lifeless. Lists of creative occupations—from architects through industrial designers to writers—rarely mention science or engineering. Design is often presumed to be individual creative

Preface

ix

emotional expression through form, appearance and symbolism. Function is regarded as prosaic and creatively inferior, and in any case delivered by the precision of science. Few people appreciate that science is incomplete, and its use must be judged carefully. Largely unappreciated is the idea that aesthetics is more than just beauty but concerns our emotional relationship with things through our creative artfulness, ingenuity and inventiveness. Fewer people still associate engineering with sustainable, resilient quality of life and well-being—indeed many blame technology for the ills of society. The voice of engineering has been weak. Assessment of the work of engineers has been left to other disciplines. They have tended to focus on technology as applied science without recognising what Aristotle called phronesis or practical wisdom—the intellectual virtue of practical reasoning. Individual perception and knowledge of engineering seems to depend on whether people have family or friends who are engineers. Research indicates that most non-engineers rightly associate engineering with construction, mechanics, building and fixing things but miss the aspects of design, innovation and creativity, problem solving and aesthetics. They see engineering as the use of a preset body of knowledge onto a problem. The media often compound the problem by rarely referring to professional engineering as creators of technology—still less as creative designers and problem solvers. Nobel was an engineer and yet there is no Nobel Prize for engineering. A USA survey showed that typically ‘in the American public mind, engineers don’t save lives…and are not involved in creating innovative materials’. The USA National Academy of Engineering launched a campaign in 2013 called Changing the Conversation. Their four messages are: engineers make a world of difference; engineers are creative problem solvers; engineers help shape the future; and engineering is essential to our health, happiness and safety. A key reason for writing this book is that the lack of public acquaintance of the role of engineering in modern life hinders the recruitment of the best of young talent —and most importantly women and black and minority ethnic groups. Engineering is perceived by many to be male-dominated and yet more and more women engineers are enjoying inspirational careers. Examples include Julia King, Baroness Brown of Cambridge, who held senior position in Rolls Royce and was Vice Chancellor of Aston University; Sarah Buck, the 88th, but first female, President of the Institution of Structural Engineers; and Dame Ann Dowling, the first female President of the Royal Academy of Engineering. They perhaps appreciate that engineering is not only about mechanics and machines but is also a people profession. I hope this book will inspire younger people to consider a career in engineering and to maintain that career after graduation. I would like to thank a number of people to whom I owe a great deal. First to Juliet Bailey of Dash+Miller, The Bristol Weaving Mill, Bristol who was so enthusiastically helpful in showing me her workshops and looms. My thanks to Christopher Simpson and to James Crowden who have allowed me to tell their personal stories. Thanks also to Tony Copping, Bob Baird, Ray Portman, Patrick Godfrey, Oksana Kasyutich and Sally Heslop, who all read parts of the drafts of the book and provided overall feedback and comments that have been enormously

x

Preface

helpful. Sally also helped particularly in ironing out gender issues, Tony with electrical and electronic matters, Bob with chemical engineering, Patrick with systems and Ray in reducing some of the overly complicated explanations in my early drafts. Thank you to Michael Berry for useful exchanges on quantum theory. Thank you to the Special Collections of the University of Bristol for allowing me to use Fig. 4.10. Thanks to John Roberts and Brighton i360 Limited for permission to use images for Fig. 9.1. Thanks to photographer John Charles Burrow (1852–1914) and the Steve Colwill Collection who own the copyright for Fig. 9.2a. I thank all my colleagues and friends, too numerous to mention, who have over the years discussed, debated and argued many of the points made in the book. Of course, what is written here is entirely my responsibility. I thank Brian Halm, Production Contact, Michael Luby, Responsible Editor at Springer who was very helpful in getting the book published. Also, at Springer I thank Ms. Shalini Monica, Project Coordinator, Mr. Vishnu Muthuswamy, Project Manager and the production team. Last and by no means least I thank my wife Karen Blockley for her unfailing love that sustains me. Bristol, UK

David Blockley

Reference World Health Organisation. (2015). http://www.who.int/gho/mortality_burden_disease/life_tables/ situation_trends/en/. Last accessed February 2019.

Praise for Creativity, Problem Solving, and Aesthetics in Engineering

“What a terrific book. It’s also beautifully written and beautifully illustrated. Reading it makes me proud to be an engineer. And Dr Blockley explains engineering and the remarkable achievements of engineers in a way that is easily understandable to all.” —Dr. Robert Langer, Institute Professor, Massachusetts Institute of Technology, USA; Winner of the 2013 Queen Elizabeth Prize for Engineering “We need philosophical engineers. David Blockley is one of very few. He is a civil engineer who thinks deeply and widely about aspects of his profession that are usually ignored in the popular stereotype: how the practical and the scientific are inevitably embedded in the needs of society and have been for millennia. The innovative organization of the book is by activities: dreaming, making, dwelling, moving, fighting … An abundance of detailed examples, presented in their historical contexts, illustrates what engineers do, and how their work is central to our culture. This book is wise, subtle, and above all deeply human.” —Prof. Sir Michael Berry, FRS, Melville Wills Professor of Physics (Emeritus), University of Bristol, UK “This book is a tour de force distilling the scholarship, deep thought, open mindedness and practice of a lifetime in a relevant and fascinating manner. In my opinion, as a doctor, the author demonstrates convincingly that, like medicine, engineering “is undertaken by people, and for people, to improve the human condition”. Given my deep and long interest in Medical Ethics, I applaud the author’s desire to connect engineering and medicine emphasising the importance of values and ethics in both disciplines. Each has much to learn from the other and this book is a wonderful primer to enable to us to do so. It deserves a wide readership.” —Gordon M. Stirrat MA, MD, FRCOG, Emeritus Professor of Obstetrics and Gynaecology, The University of Bristol; Honorary Vice-President, The Institute of Medical Ethics

xi

xii

Praise for Creativity, Problem Solving, and Aesthetics in Engineering

“Over the past two decades or so, an increasing portion of the economics’ profession has started to see itself as “engineers” rather than as “scientists”. By this it is most often meant that these economists’ job is as much about “doing” things than about “knowing”: designing markets and incentive structures in domains where economic motives and technological challenges are intertwined. This transformation of economics into a branch of engineering has often been equated with another layer of estrangement from the discipline’s moral and philosophical origins. However, in this beautifully written essay, David Blockley reminds us, first, that while often separated in modern Western thought, “knowing” and “doing” still go hand in hand, and, second, that engineering problems always involve human and moral elements that bring out crucial accountability issues for those whose daily job consists in solving them. Drawing from years of experiences in civil engineering education, the author carefully analyses the societal and environmental challenges facing engineers today and exposes the guiding principles they should keep in mind while accomplishing their missions. This is a must-read, not just for the socially conscious engineer but for all professionals (economists, managers, lawmakers, etc.) whose role in designing things affect, either positively or negatively, people’s standards of living.” —Pedro Duarte (University of São Paulo) and Yann Giraud (University of Cergy-Pontoise), Co-editors of Economics and Engineering: Institutions, Practices and Cultures (Duke University Press, 2020) “David Blockley has taken on a vast subject—pretty much the full history of the planet, along with some solar system content with plenty of thought about where we are going. I admire the breadth of the scope of the engineering he describes. I liked the way he deftly links engineering with the other disciplines and the development of the professions. Including Sir Neil Cossons’ stirring reference to the “magic” of what we engineers do was masterful. The creative process and the ‘adrenaline rush’ of excellence in design is what ‘intoxicates’ designers like me. Bravo, David, for expertly guiding us through this important topic.” —David Harvey, FIStructE, Bridge Engineer, President of the Institution of Structural Engineers 2006–07, Vancouver BC, Canada “This illuminating book is aimed at the non-technical reader who may not realise the wide role of engineers. It includes an exciting range of stories of engineering achievements. The opportunities for young people are vast ranging from engineering workers to chartered/professional engineers. As a former CEO of a UK manufacturing company, who is not technically qualified but had engineers working under my direction, I fully support David Blockley’s view that “practical people are often made to feel inferior compared to those with more theoretical knowledge”. This is a message that needs to be promoted vociferously.” —Ray Portman, former Managing Director of FBT Holdings Ltd., Birmingham, UK

Praise for Creativity, Problem Solving, and Aesthetics in Engineering

xiii

“Scientists, it is said, look at things that are and ask why? Engineers think of things that never were and ask why not? In this wide-ranging book David Blockley seeks to “change the conversation” about engineers and engineering—presenting them as making huge contributions to our civilization and lifestyles. The book is wide in scope—tackling areas from bicycles to binary devices; long in historical reach— extending from the catapult-like trebuchet to pacemakers for the heart; and layered in description—ranging from the lives of engineers, through the technical details of their artefacts, to their social settings and repercussions. All of this conveys the author’s conviction that engineering, while using the results of science whenever available and wherever possible, constitutes practical wisdom for action—or phronesis, as Aristotle called it. His grasp of detail as well as “big picture” is very impressive.” —Priyan Dias, Senior Professor in Civil Engineering, University of Moratuwa, Sri Lanka; Vice-President, National Academy of Sciences, Sri Lanka “David Blockley writes ‘we are all engineers’ and his book really brings that message to life. It deserves to be read by everyone whether technically minded or not. He includes the diverse range of engineering disciplines as well as the historical context of how “dreams have been turned into reality”. The message that first and foremost engineering is about people is loud and clear. The book is totally in accord with the vision of Engineers Australia ‘……that engineers shape the future of Australia and that creativity is fundamental to engineering’. Their magazine ‘create’ is about engineering ideas into reality to raise the profile of engineering within the community.” —Mike Fordyce, Director CROSS-AUS, Former Principal Engineer, Kellogg Brown & Root Pty Ltd., Brisbane, Australia; President of the Institution of Structural Engineers 2004–05 “This is a masterly, well written, and thoroughly entertaining tour de force. It demonstrates clearly that “engineering is done for people by people to improve the human condition”. As the author states the “image” that most people seem to have of engineering does not readily connect with their daily experience. The book seeks to bring Engineering out of its technical silo and, in large part succeeds, but there are sections (e.g. on neural nets) that require some knowledge of mathematics. All in all, the book lives up to the cover outline.” —R. L. Baird, P. Eng., Toronto, Canada “David’s book totally captures the massive impact of engineering and why engineers love what they do. It also serves as a timely reminder of how engineering can, should and often does work as a career for women and men alike.” —Michelle McDowell, MBE FREng, Chair of Civil and Structural Engineering at BDP, London, UK

xiv

Praise for Creativity, Problem Solving, and Aesthetics in Engineering

“David Blockley has produced a timely book, interesting to engineers and the general reader alike, dealing with engineering and engineers at a deeper than usual level. The ideas are wide-ranging, from the ways in which the products of engineering support us in almost everything we do to fundamentals such as aesthetics and the necessary (and not obvious) balance engineers need between intellectual objectivity and well-grounded practical wisdom. There are also useful warnings about unintended consequences and the need for even the best engineers to be aware of the wider social and environmental context of their work.” —David Elms, Emeritus Professor, University of Canterbury, New Zealand “This is a book that inspires as well as informs. It shows, with plenty of examples, that throughout history the path to innovative success is complex and uncertain. Navigating that path between danger and opportunity means that we must be able to integrate the detail into the big picture because that is where outcomes are determined by combining an understanding of people and technical behaviours. This requires the collaboration of people with diverse world views, practical experience and scientific knowledge. I recommend this book to any who want to extend and enrich their understanding of engineering.” —Patrick Godfrey, FREng, Emeritus Professor of Systems Engineering at the University of Bristol, UK; Former Director of Halcrow and Managing Director of Halcrow Offshore, UK “This is a major contribution not just to the public appreciation of engineering but to an emerging field of the philosophy of engineering. Here an engineer with a philosophical mind reflects critically on his professional activity and proposes principles for its future development that deserve attention from all of us who now live in an engineered and engineering world.” —Carl Mitcham, Professor Emeritus of Humanities, Arts, and Social Sciences, Colorado School of Mines, USA; Distinguished Professor of Philosopher of Technology Renmin University of China

Contents

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

1 1 5 6 7 8 9

2 Making . . . . . . . . . . . . . . . . . . . Craft Making . . . . . . . . . . . . . . . Art Making . . . . . . . . . . . . . . . . Functional Art Making . . . . . . . . Invention . . . . . . . . . . . . . . . . . . Engineering Making with Science Engineering and Technology . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

13 14 15 16 18 19 23

3 Grounding . . . . . . . . . . . . . . . . . . . . . . . . . . . Five Principles . . . . . . . . . . . . . . . . . . . . . . . . . The Clockwork Universe—A Discredited Myth? A Learning Odyssey . . . . . . . . . . . . . . . . . . . . . Problem Solving with Practical Wisdom . . . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

27 28 30 32 34

4 Dwelling . . . . . . . . . . . . . . . . . . Ronan Point and Grenfell Towers Shelter . . . . . . . . . . . . . . . . . . . . Theory and Practice . . . . . . . . . . Fragmentation . . . . . . . . . . . . . . Architectural and Structural Form How to Read a Structure . . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

41 41 45 48 51 53 53

1 Dreaming . . . . . . . . . . . . . . . . Ben’s Story . . . . . . . . . . . . . . . What Is Engineering? . . . . . . . . A Clarification . . . . . . . . . . . . . Why Is Engineering Important? . Who Are Our Engineers? . . . . . All Is Not Well on Planet Earth

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

xv

xvi

Contents

Wind and Vibrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Progressive Collapse and Terrorism . . . . . . . . . . . . . . . . . . . . . . . . . . . Structural Safety . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

58 60 62

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

. 69 . 69 . 70 . 71 . 72 . 74 . 86 . 86 . 88 . 94 . 98 . 100

6 Communicating . . . . . . . . . . . . Patterns . . . . . . . . . . . . . . . . . . The Syntax of Jacquard’s Loom Signalling . . . . . . . . . . . . . . . . . Michael Faraday . . . . . . . . . . . . James Clerk Maxwell . . . . . . . . Transmitting the Human Voice . Computing . . . . . . . . . . . . . . . . Semantics and Pragmatics . . . . . Perspectives . . . . . . . . . . . . . . . Passive and Active Information . Big Data Becomes Active . . . . . Artificial Intelligence . . . . . . . . Women in AI . . . . . . . . . . . . . . Getting Smaller . . . . . . . . . . . .

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

105 105 107 109 109 112 115 117 120 122 122 123 123 128 129

7 Fighting . . . . . . . . . . . . . . Literary Engineering . . . . . . War . . . . . . . . . . . . . . . . . . Weapons . . . . . . . . . . . . . . Trebuchet—Early Ingenuity . Gunpowder . . . . . . . . . . . . Samuel Colt . . . . . . . . . . . . Baron William Armstrong . . Radar . . . . . . . . . . . . . . . . . Nuclear Weapons . . . . . . . . Rockets and Missiles . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

131 131 134 135 136 137 138 138 140 141 143

5 Moving . . . . . . . . . . Edith’s Story . . . . . . Moving . . . . . . . . . . Early Land Transport Water Transport . . . . Land Transport . . . . . Engines . . . . . . . . . . Fluids and Solids . . . Heat Engines . . . . . . Turbines . . . . . . . . . . Jet Engines . . . . . . . . Moving On . . . . . . .

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

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

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

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

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

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

. . . . . . . . . . .

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

. . . . . . . . . . .

. . . . . . . . . . .

Contents

xvii

Targets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145 Defence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147 Sustainability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149 . . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

153 153 155 156 162 165 168 171 177

9 Flourishing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . West Gate Bridge . . . . . . . . . . . . . . . . . . . . . . . . . . . Will This Never Happen Again? . . . . . . . . . . . . . . . . Revisiting Patterns . . . . . . . . . . . . . . . . . . . . . . . . . . Practical Wisdom and Rigour . . . . . . . . . . . . . . . . . . Grand Challenges . . . . . . . . . . . . . . . . . . . . . . . . . . . Engineers at the Heart of Society . . . . . . . . . . . . . . . . Meeting the Challenges . . . . . . . . . . . . . . . . . . . . . . . Fragmentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Learning Together Requires Leadership . . . . . . . . . . . Creativity, Problem Solving and Aesthetics—Turning Dreams into Reality . . . . . . . . . . . . . . . . . . . . . . . . . Engineering Is a People Profession . . . . . . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

181 181 187 188 191 194 201 202 204 205

8 Well-Being . . . . . . . . Robert Langer’s Story . Well-Being . . . . . . . . . Agricultural Making . . Drinking Water . . . . . Waste . . . . . . . . . . . . Joseph Bazalgette . . . . Pacemaking . . . . . . . . What Next? . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . . . . . . 206 . . . . . . . . . . . . . 209

Glossary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213 Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219

Chapter 1

Dreaming

You can design and create and build the most wonderful place in the world. But it takes people to make the dream a reality Walt Disney

Ben’s Story Ben Telford is a 74-year-old structural engineer who owes his life to electrical engineers. Fifteen years ago, Ben (not his real name) had a heart pacemaker smaller than a matchbox implanted in his chest. Without it he would die. Before it was fitted, he was only dimly aware that such things exist. As he learned more, he became hugely impressed by the way that an assorted bunch of people—doctors, scientists and engineers, largely unknown to the rest of the world, turned a dream into a reality. By co-operating and learning together they developed a lifesaver more or less in Ben’s own lifetime—the treatment wasn’t there for his father who died of the same heart condition only 40 years ago. Ben’s story started when he and his wife Jenny were driving through a small town. He stopped to allow some pedestrians to cross the road. What followed was good and bad news. Good news because what happened could have been so much worse: if Ben had been driving on a motorway, he and his wife and others may well have been killed. Bad news because Ben suddenly felt a strange surge of pressure rising up from his chest almost like a tidal wave to his head and, for a split second, he lost consciousness. Jenny saw his distress. ‘Whatever is the matter?’ she asked. Ben’s reply was uneasy ‘I’m not sure but I think you’d better drive’. Back home Ben went to see his doctor. She was firm: ‘You must not drive a car until the problem is sorted out. I will arrange for you to see a cardiologist’. A few days later the heart specialist was listening to Ben’s heartbeat through a stethoscope and connecting Ben up to an electrocardiogram or ECG—a device that checks heart rhythms and electrical activity. He deftly placed 12 sticky electrodes © Springer Nature Switzerland AG 2020 D. Blockley, Creativity, Problem Solving, and Aesthetics in Engineering, https://doi.org/10.1007/978-3-030-38257-5_1

1

2

1 Dreaming

on Ben’s arms and legs and chest. The placing seemed a bit imprecise to Ben. Each electrode had a lead attached to a machine which looked something like a laptop computer. Afterwards Ben said it felt as though a giant dead spider had dropped on him. The cardiologist pressed a button, and the machine spewed out a long print out of squiggles on graph paper. He showed them to Ben. ‘This is a plot of your heart-beat’ he said. ‘You can see that your rate is pretty irregular. You have atrial fibrillation—a condition in which the electrical control system of your heart isn’t behaving properly. As you know the heart is a muscular pump. The electrical signals that synchronise all your muscle activity originate from the top right of your heart— your Sino-Atrial stimulus or SA Node. The role of the SA Node is to depolarise the heart muscle cells. Depolarisation just means that there is a change from the negative charge inside a cell (compared to the outside of the cell) to a positive one that allows electrical pulses to flow. As a consequence, a wave of contraction is sent through the heart. If this mechanism isn’t working properly and the gap between one heartbeat and the next gets too long, then you may feel dizzy because your brain isn’t getting enough blood. You may even ‘lose it’ for a moment until the next beat kicks in. That’s what I think happened to you when you were driving. My advice is that you need a heart pacemaker. We can arrange that within a month or so’. ‘How common is this condition?’ asked Ben, ‘and will I be able to live a normal life afterwards?’ ‘Very common’, came the reply. ‘Around one million pacemakers per year are implanted worldwide’. ‘It’s no big deal then’ said Ben. The cardiologist said ‘Almost all of the people live normal lives just as before—albeit with some medication and regular check-ups—there are increased risks, but we can manage them’. After Ben got home and had time to digest what had happened, he began to feel rather better. After all, things could have been much more serious—he and Jenny could have died, and he could have caused the death of others. He was about to join a unique but fast-growing club of special people who directly owe the quality of their lives to the newly burgeoning topic of medical engineering. But Ben’s story didn’t end there—many years later there was an unexpected twist. Ben and Jenny were on holiday and walking along the seafront when quite ‘out of the blue’ he again momentarily lost consciousness and collapsed. His heart had fibrillated. In other words, his left ventricle had, for about 10 s, gone into a very rapid heartbeat. For fifteen years his pacemaker had served him well. But in all that time there was something he hadn’t fully appreciated. His pacemaker wasn’t designed to cope with very rapid heartbeats or tachycardia—it could only prevent a slow heartbeat. Back home Ben’s cardiologist said he now needed an ICD—an implantable cardioverter-defibrillator that delivers shocks to the heart if it detects life-threatening arrhythmias. Later when he had time to reflect on this latest incident, Ben couldn’t help comparing his engineered implants with the bridges that he had designed as a structural engineer. Both work as they should within well-defined boundaries. Go outside those boundaries and they are no longer fit for purpose. For example, all road bridges have a weight limit and some bridges need constant maintenance because they are susceptible to wear and tear. Ben’s pacemaker had worked for fifteen years with no

Ben’s Story

3

tachycardia—in effect no overweight truck had come along to threaten his ‘pacemaker bridge’ and there was little wear and tear. Then one day the unexpected and unwanted happened and his pacemaker had to be replaced by an ICD. Ben felt it was as if they were dismantling one of his bridges and building a completely new and different one. On the surface Ben’s own pacemaker, and the bridges he has designed, seem to be very different—yet they are engineered products. With such diversity how do we capture what engineering is all about? What kinds of people become engineers? What are the stories behind some of the important engineers of the past such as Michael Faraday, Joseph Bazalgette and Frank Whittle as well as some of the less well-known engineers of the present? What are the different types/branches of engineering? Is engineering a suitable career for women? What does engineering do for ordinary people? In what ways do engineers contribute to human well-being? Do they have any roles in aspects of life perhaps not normally thought of as engineering such as public health, water, food, agriculture and medicine? How does engineering relate to other kinds of making, such as craft, fine art and invention? What are some of the common myths about engineering? Is engineering simply an applied science? Indeed, does science have all the answers we need? How do engineers coax the forces of nature to turn dreams into reality? Do we expect our buildings to be ‘as safe as houses’? What happens when, sometimes, they fail? Should we always be looking for someone to blame? How safe is safe enough? How did engineered artefacts such as ploughs, bicycles, engines and electrical equipment such as radios, computers and mobile/cell phones come into being? Did Ben’s heart pacemaker, evolve over time? Are the technical differences between solids and fluids important? Why are ‘patterns’ so crucial to human communications? What about the darker aspects of engineering? You will be aware that engineers have created ever more efficient ways for us to kill each other. Engineers build nuclear weapons, missiles, remotely controlled drones and enable cyberattacks on our information systems by hostile powers—including terrorists. And this darker side is not confined to warfare. Engineers are responsible for many of the things that are contributing to climate change. The computers and mobile cell phones are the carriers of social media with some unfortunate consequences such as the so-called dark web. Taking all these developments, good and bad into account what indeed are the global ‘grand challenges’ that we face together as a human race? The answers to these questions in this book will reveal some of the allure and fascination of engineering. A crucial theme that runs throughout is how we learn from what we do. Can we ensure that failures and disasters ‘never happen again’? Are there ‘patterns’ we can spot as failures incubate? Can we use those patterns to act to prevent or reduce future impacts of failure? Can we understand the risks better? How do we make our systems more resilient? If Ben were asked ‘what do engineers do?’ then he would probably answer, as would most engineers, that engineering is about solving problems. What is more different kinds of engineering require different kinds of solution—there is no onesize-fits-all. But his pacemaker/ICD experience reinforced in him the important idea

4

1 Dreaming

that all engineered products have limitations—some clearly stated (like a weight limit on a bridge) and some unstated (like tachycardia and his pacemaker) and some unforeseen (like the 9/11 terrorist attack on the WTC in New York in 2001). Engineered artefacts can only perform properly, or be ‘fit for purpose’, when designed for that specific purpose or set of purposes. It emphasised to Ben that the first creative act of the engineering process is to identify those purposes (that are necessary in a particular context) and then to have the practical foresight to imagine unknown future scenarios where unforeseen demands are placed on any solutions they may devise. In Chaps. 3 and 9, I set out five grounding principles to help tackle these issues. One of them, the principle of the unintended, emphasises practical foresight as a quality we have consistently undervalued in the past. The biggest lesson from Ben’s story, however, is one we can all relate to—it is that every one of us depends on engineering so much that we take it for granted. Ben, like millions of people around the world with heart problems, relies on a pacemaker— without it he would die. We expect the pacemaker to be safe in all circumstances—but it is designed to be safe only in a context as Ben discovered. More mundanely, every one of us crosses a bridge regularly—perhaps daily. Unstated, but expected, is the assumption that the bridge is safe and will not collapse as you cross it. However, that assumption was cruelly exposed when the Morandi Bridge in Genoa collapsed suddenly after over 50 years of service in 2018.1 There had been maintenance issues, largely unknown to much of the travelling public but known by the owners, which eventually resulted in the cables snapping suddenly killing 43 people, injuring 16 and causing massive damage to the local environment. As I wrote in the Preface engineers make buildings, roads and bridges, cars and trains, ships and aeroplanes, TVs, radios, computers, smartphones and lots of other things too—some of which we will look at in the rest of the book. We expect these things to function effectively and to be safe—and they are—that is why so many of us tend to take them for granted. But that does not mean that occasionally, like any human activity, something can’t go wrong. As we routinely use these things, we should remember that engineering is not only something-to-do-with-engines and it is not scientific certainty—we engineer when we are ingenious, inventive, resourceful and aesthetic but also when we are human. The influence of engineering goes right into the heart of our economy and wellbeing. Clem Chambers, the financial entrepreneur, wrote2 : ‘It’s technology, not politics, that creates the feel-good factor’. He continued, ‘Human wellbeing does not correlate with the presence of great leaders or polities. It correlates with technology. There have never been so many humans and they have never lived so long, and that’s because of engineering… Technology is deflationary, it makes things cheaper and better… keeps inflation down and just as it deflates prices by making things cheaper and better, it creates a kind of economic growth that does not appear in GDP but does appear in employment. Luddites say technology will make you unemployed, but in fact the opposite is true’. Notice that Chambers refers to technology not engineering—the differences are subtle but important and we shall explore them in Chap. 2. For the moment we can

Ben’s Story

5

think of them as synonymous. So now we must address our first question—just what is engineering?

What Is Engineering? Right from the start, we humans have made things—we have engineered them out of necessity to meet our needs—just like all living creatures. Ants and birds build nests for shelter, beavers build dams to create ponds to protect themselves from predators, chimpanzees use sticks to harvest honey and fish for ants, and killer whales co-operate to herd smaller fish for food. Our ancient ancestors could do much better with their big brains and consequent ingenuity—they built huts for shelter, used spears to hunt for food, and invented wheeled wagons to carry goods and ploughs to till the soil to grow food. Innovative irrigation and farming methods in settlements led to the development of cities connected by road and canal networks. Much later came the well-apportioned housing and impressive temples of the Greeks and Romans. The cathedrals of the mediaeval period were magnificent, awe-inspiring and monumental. Up to the Renaissance, there was little distinction between roles—the same person might be artist, architect, craftsman, mason or engineer. But then the jobs began to fragment as the influence of science grew. The seventeenth–eighteenth-century steam engines were more prosaic than the cathedrals but nonetheless phenomenal as they pumped water out of mines to extract the coal and iron for approaching industrial revolution. As science continued to develop so we specialised and made the engines, skyscrapers, aeroplanes, electricity, TV and computers that we know today. Only 150 years ago, Ben’s heart pacemaker wasn’t even a dream, and the first tentative attempts towards such a device were made less than a century ago. Throughout our history we have used our inventive, resourceful and skilful ability to make life a little better for ourselves—we have used our engineering ingenuity to turn our dreams into reality. Formal definitions of engineering are usually more ‘matter of fact’. Engineers think of themselves as problem solvers. Perhaps the most authoritative statements are those of the premier engineering society of the UK, the Royal Academy of Engineering. They say3 : Engineers make things, they make things work and they make things work better. Engineers use their creativity to design solutions to the world’s problems. Engineers help build the future. Engineers work in almost every area that affect people—including biomedical engineering, like new materials for hip replacements or advanced prosthetics. Engineers make the food we eat and the medicines we take. They also develop new materials like high performance sports fabrics or new electronic displays. Dictionary definitions often stress that engineering is the application of science and mathematics to solve problems. For example, the online dictionary.com says4 engineering is the art or science of making practical application of the knowledge of pure sciences, as physics or chemistry, as in the construction of engines, bridges, buildings, mines, ships, and chemical plants. Similarly, Oxford dictionaries online

6

1 Dreaming

states engineering is the branch of science and technology concerned with the design, building, and use of engines, machines, and structures. The unstated implication is that the art of engineering is a skill or technique (as an ability to apply rote scientific procedures or methods to achieve a desired outcome) rather than a creative expressionist aesthetic ability to solve a problem using science to achieve a desired outcome. The curious thing is that these definitions hardly mention people and their individual needs—strange since engineering is done by people for people to make life a little better. In that direct sense engineering is a people profession. Expanding that idea a little, my suggestion is that: Engineers produce ‘things’ that we ‘need’ like potable water and heart pacemakers and some things we don’t need, but may like to have, like a home weather station or a smartwatch. And I use the word ‘produce’ to include the many different types of creative problem solving needed to conceive, design, make and maintain—with practical foresight and wisdom.

A Clarification I need to clarify the two words in quotes—things and need—because they help us to see a wider view of engineering than is often the case. They will help us appreciate better perhaps the breadth of influence of engineering. First, I have used the rather vague word ‘thing’ and I need to explain why. Normally we use ‘thing’ to mean something we can’t or don’t want to identify precisely. A thing is an object in our minds and unclear until we give an example. It might be a material inanimate thing like a car or a bridge—objects that we usually associate with engineering. Alternatively, an object might be a material living thing such as an animal, plant or human being. That enables us to include bioengineering which ranges from making human tissue and medical devices, to new biomaterials (such as a hip implant) and biomanufacturing (such as vaccines). Finally, an object might be an immaterial and abstract idea like the number three or a belief in the afterlife or a fear of dogs. That enables us to include engineering as abstracting ideas such as a new theory or way of ‘looking at the world’. For example, the meanings of the colours of traffic lights are an abstraction we all share. As we shall see in Chap. 6, the levels of understanding of computer systems are patterns of levels of abstraction. Planning, organising or creating new ‘states-of-affairs’ are abstract activities—we might speak of engineering a good outcome—for example, engineering a good career for ourselves. Few people would think of ‘engineering’ a wedding, but if you have ever been involved in one then you may appreciate that organising it is, to a degree, like delivering an engineering project with lots of problems to solve. Organisers have to understand and distinguish between what people need and what they want. Plans have to be put in place and the actual sequence of events thought through. Everything will work better if all of the individuals and organisations work together in harmony towards a common goal. Engineering projects and weddings have the typical stages of the process of problem solving that we’ll look at in Chap. 3. So engineering is not

A Clarification

7

just about making a singular thing it is also about making and organising systems of things—from a laptop computer (see Chap. 6) to an airport, an air-traffic control system, a flood prevention scheme and the World Wide Web. The second word ‘need’ in my interpretation of engineering is also crucial because it drives the processes of how we engineer. Needs define the reasons why we are doing what we are doing. Defining the needs as clearly as possible is at the heart of finding common purpose and encouraging people to collaborate. One of the first tasks for the engineer is to distinguish what a client, customer or person needs from what he/she wants. At an individual level right from the day we are born we have needs and as expectations rise they become ‘wants’. Needs are things that are essential like food and water but also higher level things like belonging, curiosity and wonder. Wants are what we desire but are not essential like weather stations and designer clothes but also might include ambition and fame. Wants can become demands through our ‘willingness-to-pay’. You may want a luxury item but if you cannot pay for it you create no demand. Distinguishing between needs, wants and demands by managing expectations is important for our well-being since wants can easily become the unaffordable demands that lead to debt. Making that distinction entails an understanding of the aesthetic of living that is so often lacking in modern life. We all (and engineers in particular) need to be much more aware of the quality of our emotional relationship with objects including the nature and expression of beauty, style and taste that lifts our spirits and contributes to our mental health. In summary, the existing definitions of engineering written mostly by engineers do not acknowledge the connection between what engineers do as technical specialists and what they do as ordinary human beings. My more general conception incorporates the specialist definitions and helps us to see that engineering is part of being human—something we all do every day. The crucial difference is the nature of the required publicly acknowledged and specialised expertise.

Why Is Engineering Important? Sir Neil Cossons is a principal advisor to the UK Government on the historic environment. He was the first Director of the Ironbridge Gorge Museum, has been Chairman of English Heritage and Director of the Science Museum, London and Chairman of the Council of the Royal College of Art. He says5 : To engineer is human. Our capacity to make and use tools is perhaps the decisively distinguishing feature of our species, marking in the transitions from stone, to bronze, to iron, the early epochs in human history… … … …The engineering heritage matters. It should be important to engineers, as an expression of what they are and do, a part of their pedigree, and an inspiration for the future. It is important to the wider public - for the very same reasons. It needs to be seen by engineers and the wider public as an integral part of the wider heritage, indivisible from everything else that we wish to take forward with us into the future. It requires special skills and those in shortest supply are often the non-engineering skills to do with narratives and context, good visitor management and interpretation. To engineers my message is, spread the word to the rest of the world, get out of your professional bunkers

8

1 Dreaming and meet the people, ignite in them some of the magic of what it is that you do. And, join the heritage world. To the heritage sector my message is, marvel at the achievement of expert and enthusiastic engineers, recognize their qualities, help bring their preservation principles and interpretive skills up to date, and offer them a place at the table.

When you think about engineering, I wonder what comes to mind. Typical lists include engines, railways, cars, aeroplanes, buildings and bridges. But you may not have included hearing aids, the Internet or nanotechnology? One way of imagining why modern engineering is important is to think of the things you rely on most days of the week and you possibly take for granted. What would life be like if they were all suddenly taken away? Ben, of course, knows he wouldn’t be around at all without his heart pacemaker. More routinely we wake up in the morning and switch off the alarm clock—we would get rather cross if we found that the wrong time was showing. We get up and have a shower—how does the water get to the tap? We turn on an electric light and plug in the electric kettle to make a cup of tea or coffee without a thought of how the electricity got to the socket. We have a quick breakfast—of food grown with the help of tools and machinery, processed in a factory and supplied to the shops by trucks. We reach for our mobile cell phone and check text messages and emails without wondering about the highly interconnected network of computers behind it all. We travel to school, work or to meet friends in a car, by bus or train. And so the day continues. Without these modern engineered products, life would not stop but our quality of life would definitely deteriorate, be less convenient, uncomfortable and not as efficient. But more than that our daily routines and tasks would have to change completely. Collectively our productivity levels would drop dramatically.

Who Are Our Engineers? If you have ever had a technical problem at home with a home appliance like a dishwasher, central heating system, TV or computer then the repair company may well have sent out a man or woman who they described as ‘the engineer’. Clearly, some engineers look after equipment and fix problems. But, you may or may not be aware, that there are other engineers who are responsible for significant and practical complex things like big bridges, telecommunications, power stations and water supply. You may know of engineers who are ‘applied scientists’ and others who make hospital equipment like X-ray machines, CT and MRI scanners and artificial hips. What is it that professional engineers have in common? How do we go about classifying engineering activity? One way is to distinguish the different scope and responsibility of work from the specific industrial expertise required. The scope of work ranges from creating the first idea to making the final reality—from the conceptual dreaming of the visionary designer to the harsh practical manual labour required to put the idea into practice— from the idea of a building to the tasks of laying the bricks and wiring up the electrics. In the UK, four levels of scope are described as engineering worker, technician engineer, incorporated engineer and chartered engineer. Different names are used in

Who Are Our Engineers?

9

different countries, but it is common to describe them all by the one word ‘engineer’. In the USA the phrase, chartered engineer, is not used—the (not exact) equivalents used there and in many parts of the world are professional and registered engineer. These ‘vertical’ levels of expertise are important protections for public safety and in most (but not all) countries are regulated by law. They are used to define jobs, qualifications and career paths to make sure only those suitably qualified can take on particular responsibilities. Across these vertical categories, there are at minimum six horizontal divisions or categories of engineering activity. They are civil, mechanical, electrical, chemical, computing and, more recently, medical engineering. The names are reasonably selfexplanatory—except for civil. In the eighteenth century, all non-military engineering was called civil engineering. As the need for more specialist engineers developed so civil engineering has come to refer to construction and infrastructure engineering. As we shall explore later the fragmentation of the professions into these specialisms has evolved because of the increasing influence of science. Out of the first civil and non-military engineers, the specialisation of engineering the steam engine and the railways, emerged the profession of mechanical engineering. Later as our knowledge of electricity grew then there emerged the profession of electrical engineering. Later still came the chemical engineering profession, computing engineering and more recently medical engineering. But these six categories do not tell the whole story because currently in the UK there are some 36 registered professional engineering societies with an even greater proliferation of affiliate bodies. In the USA, there are even more varying from the American Society of Agricultural and Biological Engineers to the American Society of Women Engineers. As we shall see later in the book there is now a strong move to begin re-integrating the profession, and to find a common voice for the profession, as we tackle the interdisciplinary challenges of the twenty-first century.

All Is Not Well on Planet Earth Many of these challenges stem from our becoming so good at making changes and improving our lives that most climate scientists say we are close to triggering multiple tipping points—thresholds for abrupt and irreversible changes to the climate. Many of us are fearful that one of the consequences of modern technology is that we are losing aesthetic contact with things and hence control of our collective futures. Four examples illustrate the point. First, the potential for accidental self-destruction by nuclear weapons is on a scale that numbs our minds. Second, we are facing an ecological crisis. We are overexploiting natural resources, destroying biodiversity, plastic bags pollute our oceans and we face the risks of more extreme weather events as our impacts on the planet drive climate change. Third, social media, instant communications around the globe and the risks associated with intelligent robots are changing behaviours in unanticipated ways. Fourth, some of us feel, at times, that the pace of modern life has become overwhelming. For example, communications have

10

1 Dreaming

never been easier—yet many people are feeling lonelier. In 2015 the NHS choices website6 reported on a study at Brigham Young University in the USA of more than 3.4 million people. The conclusion was that social isolation is increasing the risk of premature death. One answer is to try to get away from it all—reject our new fascination with staring at screens and get ‘back to nature’. Let’s live as in the ‘good old days’ and have more time to ‘be ourselves’. ‘Being in touch with Mother Earth’ by living off the land would not completely divorce us from engineering—rather we would be back to living as people did years ago. We would still suffer from the pollutants of others and the effects of their warfare. We would still need tools, some way of making artificial heat and light, own some kind of wagon or wheeled vehicle pulled by horse or engine and, at times of great need, perhaps expect access to modern facilities like hospitals and possibly even processed food such as tea. Whilst some of the stresses of life might be relieved others would inevitably increase. Collectively we can no longer escape from the consequences of our own actions. Going back to nature may well be attractive to a few but is impractical for most of us who benefit from modern advances in living standards and health care. There are some big questions over our future collective well-being. Engineering has always been part of who we are and who we have become and therefore will be part of our future well-being. Whether we consider engineering to be the problem or part of the resolution we have a duty to future generations to ‘come to terms’ with what it is and why we need it. A good place to start is to take Neil Cossons advice and ‘take engineering out of its professional bunker….’ by positioning engineering in context. That suggestion requires us to look at the close relationships between engineering and other, perhaps more familiar, forms of making such as craft, art and technology as well as the role that science has played in increasing specialisation and the forming of professional ‘bunkers’ or ‘silos’ that often make communications between professions difficult. Then we can move on to ‘ignite… some of the magic’ advocated by Neil Cossons, some of the allure and fascination, by looking at the creative achievements and formidable future challenges of the engineering of reconnecting our aesthetics embedded in dwelling, moving, communicating, fighting and well-being. End Notes 1. The Ponte Morandi in Genoa Italy was opened in 1967 and operated as a road bridge over the River Polcevera until 2018 when it collapsed suddenly. See https://en.wikipedia.org/wiki/Ponte_Morandi. 2. See https://eandt.theiet.org/content/articles/2019/06/money-markets-it-stechnology-not-politics-that-creates-the-feel-good-factor/ (last accessed October 2019). 3. For a modern video introduction to engineering see ‘This is Engineering’ at https://www.thisisengineering.org.uk/ (last accessed January 2020) 4. See http://www.dictionary.com/browse/engineering?s=t (Last accessed February 2019).

All Is Not Well on Planet Earth

11

5. Cossons, N. (2009) Does the Engineering Heritage Matter? 3rd Australasian Engineering Heritage Conference, Dunedin, New Zealand. 6. See https://www.nhs.uk/news/mental-health/loneliness-increases-risk-ofpremature-death/ (Last accessed February 2019).

Chapter 2

Making

Think both inside and outside of the box

Making fills every aspect of our lives. The products on the supermarket shelves, the bestselling books in stores, the art in museums, the cars on roads and the search engine Google were all made by people. We need, therefore, to probe how different kinds of making such as craft, art and invention relate to each other and to engineering. An important first question has to be ‘Is there something inherently different about people who are makers?’ Steve Jobs, co-founder of Apple, didn’t think so. He is quoted as saying: Everything around you that you call life was made up by people that were no smarter than you and you can change it, you can influence it, and you can build your own things that other people can use. Once you learn that, you’ll never be the same again.

He was right but of course Steve Jobs was a special talent—a businessman, engineer, inventor and designer. But he didn’t work alone. His success depended on working closely with blending talents such as Steve Wozniak, the electronics engineer and developer, in 1976, of the first Apple computer. And we have to recognise that we each have different abilities—but the fact is that you don’t need permission to create. You don’t need to be gifted or brilliant. You just need to choose to do it and to recognise your interdependence with others—in other words, teamwork—and be much more aware of our relationship with the things we rely on. The people and the teams who choose to be makers create the environment that the rest of us live in. Michèle Dix agrees but particularly emphasises the important role that women can play. She says that recruiting more women into engineering is key to unlocking billions of pounds of economic growth. She should know—after a career in transport planning, she is now the Managing Director, Crossrail 2 in London. The first Crossrail due in 2021 is Europe’s largest infrastructure project—a 118 km railway line running through parts of London and Berkshire, Buckinghamshire and Essex. Crossrail 2 is © Springer Nature Switzerland AG 2020 D. Blockley, Creativity, Problem Solving, and Aesthetics in Engineering, https://doi.org/10.1007/978-3-030-38257-5_2

13

14

2 Making

another massive £30–40bn railway scheme proposed to open in 2033. The line will run from Wimbledon in the south west of London to Tottenham Hale in the northeast with links to major infrastructure hubs such as Clapham Junction, Victoria and Euston. Michèle has a big ‘making’ job but, like many engineers, she started off simply. She graduated in civil engineering and carried on to research for her Ph.D. in 1982 looking at the relationship between land use and transport. She then joined the Greater London Council’s graduate training scheme and became a Chartered Civil Engineer. After six years she moved to Halcrow (consulting engineers) and during her 15 years stay was promoted to be their board director for urban transport planning. In 2000, she joined TfL (Transport for London) as co-director of congestion charging on a job-share arrangement. She developed London’s Low Emission Zone and went on to become Managing Director of Planning in 2007 when she led TfL’s strategic thinking on future transport needs.

Craft Making Many of us, like Michèle I am told, enjoy baking cakes. Others like cooking, gardening or taking photographs. People love to create delightful and interesting things for a variety of reasons—enjoyment, fun or self-expression, to feel valued, fulfilled, to make something beautiful or to meet a practical need, to learn, to solve a problem or simply as a job. These are all aspects of the aesthetic of living. Engineering craft skills include bricklaying, plumbing, electrical wiring, woodworking, welding and steel fixing. These are all creative jobs for craftsmen and apprentices, turning the design choices of others into reality. But those jobs are hardly ever appreciated as aesthetic because good quality work as creative artfulness done with pride is rarely acknowledged although the results of that work are critically important. In his book ‘The Craftsman’ Richard Sennett1 defines a craft as an enduring human impulse, the desire to do a job well for its own sake. In that sense his craftsman is an engineer and encompassing all human activity—art, traditional crafts, trades and even making (wedding) arrangements. It is a unifying aesthetic. Sennett includes professionals such as doctors, lawyers and engineers and more widely, for example, the skills of parenting and citizenship. He says that every good craftsman conducts a dialogue between specific practices and thinking which evolves into sustainable habits to establish a rhythm between problem solving and problem finding. But Richard Sennett does not address the important reason why, since the Renaissance, engineering making has become separated from craft making—through the explicit and extensive use of science, nor does he seem to recognise the implications of the newly found complexities of the twenty-first century. Of course, a craft requires particular skills and knowledge—from experience but not necessarily from science. Hobbyists usually work on a small scale and often join groups of like-minded people. Crafting is not just the province of a talented, ‘inspirational’ few—though they may be particularly good at it. The joy of being able to make something that comes from yourself not only engages you more, but

Craft Making

15

it helps you to learn about who you are. Making things helps us to find meaning in our existence—it speaks to us without words. In his book ‘Why we make things and why it matters’, Peter Korn2 writes about his work as a furniture-maker. He talks about the three goals expressed in his work—integrity, simplicity and grace. In the first century AD, the Roman architect/engineer Vitruvius articulated a similar set of objectives—firmness (integrity or soundness of form), function (looking for simplicity in complexity), and delight (grace as elegance of form) that inspire many architects and engineers. These are a set of aesthetic rules for practice. Peter Korn says that to reach his goals—to create something good—one must know something good. By that he means of high quality both morally and practically. Every man-made thing, be it a chair, a text, a photograph or a school building, is a thought-made substance—although complex things like buildings are the results of thoughts of teams of people. A crafted object is the disclosure of someone’s ideas and beliefs. Richard Sennett and Peter Korn agree that we can achieve a more humane material life, if only we understand better the making of things.

Art Making Understanding and creating a work of art is perhaps more complex—and certainly a constant topic of discussion. It is in art that we usually find the word aesthetic—to mean the nature and expression of beauty of form, colour and sound. We usually know what we like when we see it but are unsure of the value of the art we do not like—such as, in my case, a surrealist painting like Edvard Munch’s ‘The Scream’. A perennial question is what is good and bad art? For an answer, we tend to rely on those who profess to know. As Grayson Perry said in his 2013 BBC Reith Lectures ‘Who validates (art)? A cast of characters…. in this validation chorus…. artists, teachers, dealers, collectors, critics, curators, the media, even the public maybe…they form…. this lovely consensus around what is good art’. Art objects speak to our emotions in ways that functional objects do not. We feel that art objects are unique—the work of a single talented and inspired individual. They have rarity value. We really don’t know how the consensus that Grayson Perry referred to is created. But we usually assume that the artist has some kind of expressive purpose—to communicate a message, however subliminal, intuitive or enigmatic. The message may be a feeling, an emotion or revelation of sheer beauty, a religious act of devotion or a political or social comment. Art illuminates us and enriches our emotional experiences. Art impacts on our social well-being, cohesion, our health, our educational system and our economy. Many different forms and structures of matter are used such paint on canvas, carving of wood or stone or even a large metal structure like the Kelpies in Scotland (Fig. 2.1a). (A kelpie is a mythical water horse that enticed people to ride on its back, and then took them to a watery grave). Artists generally rely on traditional craft techniques such as paint, wood and stone working tools and pottery wheels to turn their ideas into a reality. The Kelpies were designed

16

2 Making

Fig. 2.1 a The Kelpies, Scotland 30 m high stainless steel sculptures guarding a new extension to the Forth and Clyde Canal. Image by Beninjam200, CC BY-SA-4.0 via Wikipedia Commons. b Her Secret Is Patience designed by Janet Echelman in Phoenix, Arizona. Image by Stuart Peckham. CC-BY-SA-2.0 via Wikipedia Commons

and built through some quite difficult engineering because each structure is very large. The internal forces induced by the sheer self-weight, the wind and rain are significant—no one would want to see the structures blown over. The actual shapes are complicated to build on such a scale and so computer models with advanced surveying techniques were used to generate and build the ‘horse-like’ forms—point by point on a massive three-dimensional grid. A contrasting example (Fig. 2.1b) is ‘Her Secret is Patience’ a 45 m tall aerial sculpture in Phoenix, Arizona. Artists and engineers collaborated to create a large three-dimensional multi-layered net hovering high in the air with changing colours gradually changing through the seasons. An ethereal towering but soft structure, fixed in constant motion dancing gently in the desert winds.

Functional Art Making The ‘Kelpies’ and ‘Her Secret is Patience’ are almost functional art in that they have become tourist attractions. Functional art overlaps with craft because of the dual purpose—art and utility—the boundaries are never clear cut. The furniture of Peter Korn is another example. A more straightforward example is a friend who displays an ornamental tea set in the window of her home—not to be used to make tea, but as a pleasing object that emanates a warm feeling of ‘homeliness’. A controversial example is that of an artist who claims that a urinal made by his own hand is an art object. Whether it actually succeeds as art depends on whether Grayson Perry’s consensus is reached.

Functional Art Making

17

Functional form is necessary but not sufficient for functional objects. Designers can attempt to make them art objects with elegance and style. Think of the sleek beauty of a Ferrari or Aston Martin, the sophistication of an iPhone, the awe of a cathedral and the impressive grandeur of the Millau Viaduct in France or the Salginatobel Bridge in Switzerland (Fig. 2.2). Industrial designers and architects design objects to meet human needs but usually (but not always) need engineering or scientific help to make them functional. Some objects are so straightforward that anyone can ‘read’ them easily and hence they are effectively created in the same way as art objects. By ‘reading’ an object I mean observing, understanding and interpreting it to give meaning. We often choose kitchen equipment such as kettles, cups and saucers for their appearance and we assume without much thought that they are functionally effective. Tools such as hammers, chisels and saws do vary in quality and can be made to look more or less pleasing. Other functional objects such as a plough or bicycle may be more complicated than a cup and saucer or a hammer and they are not usually considered as art but at least we can appreciate reasonably easily how they work. Some develop a passion for cars (so-called petrolheads) and develop a consequent facility to read the ‘language’ of a car and its engineering elegance or otherwise. Functional objects like Ben’s pacemaker are usually not the work of a single gifted artist but born out of the creative work of many talented individuals over a long period. They are also produced in numbers which although similar, can make

Fig. 2.2 The Salginatobel bridge, Switzerland

18

2 Making

our lives so much better. Grayson Perry’s consensus may confer an art status on the urinal mentioned earlier but the commonplace flush toilet is not normally thought of this way. But the flush toilet did not just appear suddenly one day—the modern products evolved from the first idea by Sir John Harrington in 1596. Each object may not be the unique imaginative work of an individual but is the inspired work of a group of disparate creative people who worked over a period of time to make life better for the rest of us. To really appreciate these kinds of functional objects we have to understand their provenance—how they came to be. But we will also benefit if we can appreciate an elegance of function through understanding the way potentials create the rhythms of energy flows through the object—something we scrutinise in later chapters.

Invention Inventors are special people—sometimes technically qualified and sometimes not. An invention is a novel useful non-obvious creative idea—workable at an affordable cost to meet a need. Invention is innovative and creative design where ideas from very separate areas are brought together. The borderline between invention, industrial design and engineering are not sharp. For example, in 1861 during the American Civil War, Richard Gatling got the idea of a rapid-firing gun from watching his seed planter. When tailor Jacob Davis met Levi Strauss, an importer of jean, a cotton corduroy from Genoa, Italy, in 1873, the history of casual clothing changed forever. The clockwork radio, invented by Englishman Trevor Baylis, puts together two previously disparate ideas to great effect. It has a wind-up crank to create potential energy in a spring to drive an electrical generator—an idea which eventually needed input from engineers to make it work effectively by including a battery charged by the spring. The innovative vacuum cleaners designed by James Dyson in the 1970s could be equally correctly called invention or creative engineering design. The vacuum cleaner was well known but Dyson’s clever innovation was to use cyclonic separation to remove particulates from a stream of air without any filters. The air is rotated in a cylinder so that the larger particles with greater inertia fall out to be collected. John Taylor is a professor of innovation with over 150 patents to his name. He calls himself an inventor, but he is also an engineer and businessman who made a considerable fortune by inventing a bimetallic switch for electric kettles—again a clever combining of ideas. The well-known London Millennium landmark ‘The London Eye’ was an innovative idea from architects Julia Barfield and David Marks but one which required considerable engineering expertise to design, build and now to maintain.

Engineering Making with Science

19

Engineering Making with Science Modern engineering has emerged from the separation of craft and science and the consequent specialisation needed to cope with our increasing scientific knowledge. When anyone makes anything, they naturally want to use the best appropriate knowledge available to them. Before Isaac Newton published his laws of mechanics in 1687 the only real theory available to practitioners (engineers, architects, masons and craftsmen) was geometry. Now all the powers of modern science are to hand. But to use any science to solve a practical problem then a judgement has to be made about its dependability for that purpose. The relationship between engineering and science is one of mutual support—as we do more then we know more and hence can do more in an ever-upward spiral. But we need to note that engineering and science have different purposes—in the simplest of terms engineering is about ‘doing’ and science is about ‘knowing’. Unfortunately, we can’t just apply science without understanding the context in which it does and doesn’t work. It follows that engineering is not simply ‘applied’ science rather engineers apply or use science to make decisions and judgements about what needs to be done. Science is powerful because it brings understanding and consistent explanations. Writer Siri Hustvedt puts her finger on the pulse when she acutely observed3 ‘…in popular culture science is often perceived as monolithic…while everyone… is aware that ‘scientists’ come from many disciplines and frequently change their minds…there is a powerful sense in which that they, the scientists, are on an inexorable march forward…. as it (science) methodically uncovers the secrets of nature…. The reason…. is not obscure…we…are awestruck witnesses to technological changes that are nothing if not a testament to scientific research and its practical application’. She continues—the reality of the relationship between our scientific theories and the natural world is undeniable but it has had a blinding effect on many. She makes a strong point. In a moment I will call that blinding effect a failure to appreciate the as-is distinction—treating a theory as if it is true does not logically imply that it actually is true. Understanding context is key. All scientific theories are underpinned by assumptions about the nature of reality—and all theories are incomplete. The reason is both simple and subtle at the same time. Theory is developed and only works in a specific context—and sometimes that context is not stated as clearly and explicitly as it needs to be. History tells us that as science has developed so the context has widened—but it is never total and absolute. We cannot know the ‘mind of God’. A theory only gives way to a deeper or wider one when the domain of applicability is found to be limited. So, at any point in time, a theory is an approximation to the realm for which it was developed. For example, Newton’s laws were and still are a very good approximation for objects on the earth’s surface travelling at a small fraction of the speed of light. They have now been replaced by Einstein’s relativity theory for objects travelling at a significant fraction of the speed of light. Nevertheless, Newton’s Laws are good enough to build earthbound bridges and buildings.

20

2 Making

When a theory is applied to a practical problem then the fit is never exact— assumptions have to be made and uncertainties abound. For example, what are the weights of the heaviest trucks on a bridge and how will they change in the future? How will the strength of the steel of the core of a nuclear reactor deteriorate over its lifetime? Science does not cover everything in perfect detail that the engineer has to consider. That practical technical success proves the truth of science is a myth and misleading. There is a widely quoted aphorism ‘Scientists discover what is; engineers build what never was’. This is way off the mark. Science is not what ‘is’ but rather part of the best tested evidence we have about what we think is ‘is’ in a context. Engineers cannot just rely on the uncovered inviolable secrets of true nature. Rather engineers depend on science within a specific domain of application. They build appropriate models of nature to help them to make the decisions they need to take in order to fulfil their purpose. That purpose is to meet a human need and the models they use to do that are necessarily approximate and contextual. Engineers succeed because when they have to approximate—they make assumptions that are conservative—always erring on the safe side when in doubt to make sure they stay in context. The word model is important and controversial especially if we use it to describe scientific theory. There is little dispute when we refer to physical models, such as a toy model car or an architectural scale representation of a new sports stadium. But using the word to describe a theory is more contentious. A model is a representation of something and not actually that thing, so there are inevitably some attributes that are not included—a toy car may not have a toy engine representing the actual engine. Mathematical calculations and computer programmes are models because they represent an aspect of reality—but some attributes of that reality are inevitably not included, either because they are known but not considered relevant or because they are unknown unknowns. Engineers create models for a specific purpose and what really matters is the quality of that purpose and of the model to deliver that purpose. If essential attributes are missing, then the both the result and the model will be poor. For example, if the requirements for a new bridge are so financially stringent that only a very basic and perhaps rather ugly structure can result. When an engineer uses Newton’s laws to represent what will happen when someone drags a heavy piece of equipment across a factory floor then she makes a specific model of the interaction between the equipment and the floor within the general model provided by Newton. The floor could be smoothly polished or rough-grained, so she estimates the frictional resistance or measure it directly by tests. A river engineer uses Newton’s laws to model the volume of the flow of water over a weir. She relies on a factor called a discharge coefficient which has to be estimated from values found from laboratory tests. A structural engineer uses Newton’s laws to calculate the forces induced in the steel beams and columns of a building and in doing so commonly assumes, in his theoretical model, that welded and bolted steel joints are either pinned or fixed. In a real building, the joints will be somewhere in between—he makes his assumption to make the calculation tractable (A pinned joint between two steel structural members

Engineering Making with Science

21

means that the members are free to rotate relative to each other. If the joint is fixed, then the members rotate together with no relative rotation). After engineers have created a model of a real thing then they may apparently assume that certain values of parameters and relationships within the model are true. For example, that the joints in a steel building are pinned or fixed. However, most engineers know that assuming those values to be true in the model is not the same as them being actually true in reality. I call this the as-is distinction rather like the distinction between is and ought famously put forward by David Hume, the Scottish philosopher. In 1739 he pointed out that many people make false claims about what ought to be the case based on what is the case. In the as-is distinction, people may forget or not realise that a theory (model) successfully used as if it is true in the messy world of practical reality does not logically imply that it actually is true. The reason is that all sorts of uncertainties and fail-safe design and analytical assumptions are embedded into the context of the model. The model is not built to represent the truth of the reality but rather to serve a different purpose in that reality (such as the safety of a floor carrying the weight of the tall shelves of books in a library). Typically, the engineer will use his theoretical models to determine bounds on possible solutions. That practical success implies theoretical truth is a failure to appreciate the uncertainty assumptions ingrained into the as-is distinction and the essential difference in purpose between scientific and engineering endeavours. The success of technology does not confirm or deny the absolute truth of the science on which it is based—but it does provide us with confidence and corroboration. It helps to understand that if we do the same thing again in the same context then it will probably work. If the context changes, such as happened at Ronan Point (Chap. 4) when the industrial building system was to be built taller than anything before, then we must be extra careful to ‘dot our i’s and cross our t’s’. If we don’t do that and we are a little ‘slipshod’ then we may have an engineering accident or disaster on our hands. What we thought was the case is no longer—we have punched through our safe contextual boundaries and possibly contributed towards a set of conditions that could incubate to failure if left unattended. There are many engineering examples that stretch the boundaries successfully—Telford’s Menai Bridge, the NASA space programme, and the engineering of the London Eye. But in every case attention to detail and care in making assumptions and testing out all decisions was carefully done—although even then the projects were not entirely free of serious problems. When things go wrong (and especially if people are killed) most of us look for someone to blame—someone must be at fault. History tells us that complex projects may not always work as we intended—and we shall see that sometimes no one individual is at fault. Our widely held ‘blame’ culture derives from the idea that if uncertainty is low and that we can therefore control everything then if something goes wrong, it must follow that someone did something wrong—and worse still they didn’t care. Of course, that is not to deny that individual or corporate culpability through negligence or corruption occurs must be rooted out—but it does make us face up to the question—what do we do if there is no one to blame? History tells us that sometimes, even when people take great care, consequences happen that we only understand with hindsight—and no one was at fault. Unfortunately, in a

22

2 Making

‘blame’ culture, legal processes often hinder learning from hindsight as people are careful not to put themselves at risk of prosecution—warranted or not. If we are to find ways of being ‘fleet of foot’, avoid future unnecessary human suffering, and make ourselves adaptable to the uncertainties ahead then we have an obligation to be socially ingenious to reduce drastically the time spent on finding someone to blame. Thomas Homer-Dixon4 values ingenuity highly and says that social ingenuity is a prerequisite for technical ingenuity. So if, like him, we see our social and physical worlds as quite separate then we ought to encourage and nurture a better appreciation of, and facility for, ingenuity (and hence engineering) amongst politicians, economists and social scientists, and within our systems for health, criminal justice and our collective responses to pollution, climate change and social media. Although Thomas Homer-Dixon clearly wants to integrate the way we manage our social and physical worlds, he seems not to appreciate that engineering projects are as social as they are physical. Social ingenuity is needed ‘in spades’ for NASA’s space programme or the construction of Crossrail—something often overlooked by engineers themselves and given too little emphasis in engineering education. Projects that go over budget and overtime are not disastrous in the sense that people are killed or injured but they can be a serious waste of resources. Often, they occur because the product cycle, which as we identify later, is used inappropriately for complex projects. Examples are defence contracts for new military aeroplanes, large infrastructure projects like railways and new computer systems for large organisations like the NHS. All too often they are controversial failures. Major complex systems cannot be procured as though they are simple ‘off the shelf’ products like kettles. That is because at each stage of the problems solving processes there are considerable uncertainties. We need to think about them differently. I will argue that we have to ‘learn to learn’ our way through them. One recent example where this was done successfully was the London 2012 Olympic Games. There everyone in the supply chain worked collaboratively and the learning process was an integral part of the whole life cycle with strong leadership from Sir John Armitt. So even though unexpected and unintended consequences arose (not the least of which was the effects of the financial crisis of 2008) they were managed well. John Armitt is another example of an engineer working up from a modest starting point. He graduated with a civil engineering degree in 1966 and went to work for John Laing Construction. He spent 27 years with them working on many projects including the building of Sizewell B nuclear power station. He showed his leadership skills as project manager on the Second Severn Crossing—a major cable-stayed bridge just north of Bristol, UK. He was appointed Chief Executive of the group responsible for the Channel Tunnel high-speed rail link and then for the Costain Group which under his guidance moved from loss to profit. Then he was CEO for Railtrack and Network Rail. Perhaps his most notable accomplishment was to chair the Olympic Delivery Authority and London 2012—a remarkable success of a very complex project that not only delivered the Olympic Park but regenerated the Lower Lea Valley—then a deprived part of East London.5 40,000 people worked on the project involving thousands of businesses.

Engineering and Technology

23

Engineering and Technology In public discourse, the words technology and engineering are used almost interchangeably. The different usage inside and outside of the engineering professions are contrastingly and differently interesting. Engineers see technology as part of engineering whilst the rest see engineering as part of technology. The American National Society of Professional Engineers (NSPE) says,6 ‘The distinction between engineering and engineering technology emanates primarily from differences in their educational programs. Engineering programs are geared toward development of conceptual skills and consist of a sequence of engineering fundamentals and design courses, built on a foundation of complex mathematics and science courses. Engineering technology programs are oriented toward application, and provide their students introductory mathematics and science courses, and only a qualitative introduction to engineering fundamentals. Thus, engineering programs provide their graduates a breadth and depth of knowledge that allows them to function as designers. Engineering technology programs prepare their graduates to apply others’ designs’. As we saw in Chap. 1, in the UK, the technical professions of engineering define four kinds of qualification. In brief, they are engineering workers who are the skilled craftsmen such as bricklayers, carpenters, steel fixers, etc. Technician engineers use known techniques in a well-defined area of work. Incorporated engineers work within an existing technology but with wider responsibilities. Chartered engineers have the widest scope of work that requires creative independence of mind as designers and makers. Outside the UK the terms chartered, and incorporated engineer are not generally used but covered by the term professional engineer—someone registered to practice in a particular country, state or province. Every engineer is vitally dependent on others—mutual respect is key. Each type of engineer has an appropriate scope of work for which they have special knowledge, understanding and creative skills within the work of others. For example, a technician engineer who comes to fix your broken central heating boiler has probably not been part of its design or manufacture, but he is trained to understand how it works and to think creatively about what may be wrong and how to fix it. Craftsmen, technologists, technicians and technical engineers work within the wider remit of the chartered or professional engineers. For example, a technician engineer may design the details of a steel bolted joint within tight parameters set by a chartered structural engineer. But a chartered engineer is unlikely to be able to fix the reinforcing bars for a concrete slab or operate a crane. Each type has total respect for the contribution of others. So chartered engineers do not see craftsmen, technicians and technologists as less important—quite the contrary their work is highly valued. Nonetheless, craft and technology are more limited in scope and levels of responsibility than engineering. The relationships are rather like those of doctors and nurses—totally interdependent team working but with clearly separate roles, one primarily medical and the other chiefly caring and support.

24

2 Making

The contrast with attitudes and opinions held by those outside of engineering could not be stronger. First, the word science is typically used, especially by the media, as an umbrella word for the STEMM (Science, Technology, Engineering, Mathematics and Medicine) subjects.7 Science is regarded as the primary discipline. Technology is its handmaiden. According to this view, technology draws from engineering, the industrial arts and applied science. Increasingly and confusingly the word technology is now often used only to refer to digital information technology (IT). An example of the clash between these two views of engineering as part of technology or technology as part of engineering and both as the handmaiden of science, is the reporting of the landing of Philae on Comet 67P from the Rosetta space probe in 2014. Although widely reported as a scientific achievement, engineers landed Philae so that scientists could do their experiments exploring the origins of the solar system.8 In this chapter, we have explored some of the different ways in which we humans create and make things. Craft is a special skill, art speaks to our emotions, functional art has a dual purpose and an invention is a novel, useful non-obvious idea. Engineering has emerged from the separation of art and craft from science and the consequent specialisation needed to cope with our increasing knowledge. Engineering and technology are terms that have become confused. The view from inside the profession is that technology is part of engineering whereas, from outside, the opposite is the case. Technology is associated, in the minds of many, with the application of ‘known’ science through mere technique. Of course, engineers do apply ‘known’ science because it is highly tested and the best knowledge we have. But science is dependable only in context and does not represent the absolute truth. In other words, ‘known’ science is incomplete so that engineering has to be much more than rote application. Engineers create theoretical models for a specific purpose and what really matters is the quality of that purpose and of the model to deliver that purpose. A model used as if it is true in the messy world of practical reality does not logically imply that it actually is true. The reason is that all sorts of uncertainties and fail-safe design and analytical assumptions are embedded into the context of the model. Tackling the considerable challenges ahead in the twenty-first century will require us think ingeniously anew about the relationship between how we think and what we do to act, create and make. But how do we ‘ground’ our collective understanding of the past to provide some firmer foundations for the future? That is what we must turn to next. End Notes 1. Sennett, R. (2009), The Craftsman, Penguin Books, London. 2. Korn, P. (2017), Why we make things and why it matters, Vintage, UK. 3. Hustvedt, S. (2016), A woman looking at men looking at women, Simon & Schuster, New York. 4. In his book ‘The Ingenuity Gap’ Thomas Homer-Dixon defines ingenuity as ‘instructions that tell us how to arrange the constituent parts of our social and physical worlds in ways that help us achieve our goals’. See https://homerdixon.com/wp-content/uploads/2017/05/HomerDixon-The-Ingenuity-Gap-1995.pdf (Last accessed February 2019).

Engineering and Technology

25

5. Olympic Delivery Authority (2011).Lessons learned from the London 2102 construction project See http://learninglegacy.independent.gov.uk/documents/pdfs/ supporting-documents/learning-legacy.pdf (Last accessed February 2019). 6. National Society of Professional Engineers (NSPE) summarizes the distinction as being that engineers are trained more with conceptual skills to “function as designers,” while technologists “apply others’ designs.”. See https://web.archive.org/web/20110316114022/http://www. nspe.org/GovernmentRelations/TakeAction/IssueBriefs/ib_eng_tech.html (Last accessed February 2019). The Accreditation Board for Engineering and Technology (ABET) says that ‘Engineering programs often focus on theory and conceptual design, while engineering technology programs usually focus on application and implementation.’ https://www.abet.org/accreditation/what-isaccreditation/what-programs-does-abet-accredit/ (Last accessed February 2019). Another example is that of Century Technical College in Minnesota which says on its website ‘Engineering technologists apply engineering and scientific knowledge combined with technical skills to support engineering activities…. They are somewhat less theoretical and mathematically oriented than their engineering counterparts. They typically concentrate their activities on the applied design, using current engineering practice.’ https://century.custhelp.com/app/answers/ detail/a_id/220 (Last accessed February 2019). 7. Public discussion of these topics is usually based on the so-called STEM subjects—Science, Technology, Engineering and Mathematics. In common with some others and because of Ben’s story and his cardiac consultant’s use of medical electronics I have added Medicine as an extra M to make STEMM. 8. Let’s hear it for the Rosetta engineers. See https://www.theguardian.com/ theguardian/2014/nov/18/rosetta-all-blacks-hungary-gin-blackberries (Last accessed February 2019).

Chapter 3

Grounding

We are of this world not in this world

If the science of climate change is right (and it is the best we have), we face considerable uncertainty ahead. As we squabble about the causes, newly unfolding challenges are lying in wait. Perhaps you feel, as I sometimes do, that we are disputing the positions of the deckchairs on the Titanic to watch the looming cataclysm. The Secretary-General of the UN António Guterres said in April 2018: ‘The headlines are naturally dominated by the escalation of tensions and conflicts, or high-level political events. But the truth is that the most systemic threat to humankind remains climate change and I believe it is my duty to remind it to the whole of the international community. And indeed, information released in recent days by the World Meteorological Organization, the World Bank and the International Energy Agency shows the relentless pace of climate change’.1 In 2019, school strikes for climate action and ‘Extinction Rebellion’, with 16year-old Greta Thunberg as the standard-bearer, have led international protests with non-violent civil disobedience such blocking bridges to traffic in London. Greta has accused political leaders of not listening to the scientists. She says we need a new political system that is not about winning a competition but about cooperating and sharing resources in a fairway. The warning is serious, and the challenges are formidable. But challenges create opportunities if we can learn to recognise them. One such is to connect the new green deal with renewed social contracts and justice in politics. Unfortunately, long-term political policymaking does not fit well with short-term electoral cycles. Although late in the journey to 2050 the world then need not be a dystopia. To achieve this, I submit that we need to ground all our thinking (social, political, financial and technological) around some guiding principles—ones that are implicit in the best forms of engineering practice. © Springer Nature Switzerland AG 2020 D. Blockley, Creativity, Problem Solving, and Aesthetics in Engineering, https://doi.org/10.1007/978-3-030-38257-5_3

27

28

3 Grounding

Later in the chapter we will question some of the common myths about engineering. Does science have all the answers we need—in an age in which we have collectively begun to understand that the world is a much more complex place then ever we previously thought? What is ‘practical wisdom’ and how is it different from ‘academic’ cleverness? Is the modern distinction in education between the academic and the vocational helpful? We ask how they are similar and how they are different.

Five Principles I will start by suggesting five grounding principles that have been, in my view, implicit in the best engineering of the past and missing in the worst, and from which we would benefit if they should become explicit in the future. In brief, they are principles of Part, Unintended, Preparedness, Ingenuity and Learning. First, the principle of Part says, ‘be aware that we are of this world not just in the world and separate from it’. Our future is intimately bound up with nature. Unless we work with nature and cooperate on world scale to create a sustainable future, we will find that the future for our children will be seriously diminished. The principle reminds us that we are both parts and wholes. We are parts in the sense we belong to family and social groupings as well as the natural world. We are whole in the sense that each one of us is an individual with our own quirks and character traits. This principle helps us to harmonise the way we think of ourselves in relation to others and to the natural world. Through it we can cope better with complexity because rather than face a daunting mysterious totality we can envisage our place in the universe in layers from sub-atomic particles to the entire cosmos with each layer emerging from the layers below. We elicit this at work in later chapters. Second, the principle of the Unintended says ‘remember that everything we decide and act on has unintended consequences—some good and some bad’. The good ones are unexpected opportunities—the bad ones often can be surprise disasters. It helps us to deal with unknown unknowns—things we haven’t foreseen because we don’t know that we don’t know all that might happen. The principle reminds us that sometimes things go wrong that are no-one’s fault—we have an obligation to not immediately jump to the conclusion that someone individual is to blame. Nature can create its own surprises too. For example, when an aeroplane cannot land at an airport it may not be the airline or airport’s fault—simply the weather. The principle reminds us that contingency planning is an important part of finding a way forward as we try to cope with the unexpected and unintended events of the future. Again, the history illustrates graphically the importance of this principle. Third, the principle of Preparedness says ‘be prepared by helping to develop resilient and agile systems so we can better take advantage of new opportunities and adapt to deal with unforeseen and unintended consequences of our actions’. The principle also calls for us to change our current focus on the chance of a predicted event, to thinking through how we manage outcomes no matter how unlikely. Of course, both are important but in the past we have tended to discount important low chance

Five Principles

29

possibilities. That can make us vulnerable to small remote changes that result in very large consequences and particularly to events we haven’t predicted. Preparedness is one stage of the decision and action processes of solving problems through imaginative foresight as well as retrospective hindsight. It reminds us that robustness and resilience are important qualities of sustainability. The principle embodies some of the common sense we have undervalued through our overconfidence in the apparent certainties of science resulting in misplaced ‘triumphalism’ of human technological progress, the stress on prediction and loss of the practical wisdom of contingency planning. Fourth, the principle of Ingenuity follows. It says ‘value, nurture and develop practical wisdom’. In other words, cherish the thinking and practical inventive and resourceful skills of what Aristotle called phronesis.2 Reduce the chasm between academic and vocational education. Close the ‘ingenuity gap’ that Thomas HomerDixon3 sees between those who adapt well to complex changes and those that don’t. Move from authoritative top-down command and control to a mature collaborative culture that focuses on enabling others to be successful. The principle will remind us of the considerable ingenuity we have shown in the past but also the danger of complacency and arrogance of ‘technical triumphalism’ in the face of the challenges ahead. It reminds us that ingenuity is a capacity to be revalued, retained and nurtured. As we will trace in later chapters, it is at the core of all the best engineering and other forms of creativity. Fifth, the principle of Learning says ‘learn from our mistakes’ and from our successes. Create legal and political systems that stimulate and encourage, and do not hinder, our collective learning. It is a principle that is easy to say but very hard to do. How often do we hear the words ‘this must never happen again’ and yet history all too often continues to repeat itself? The fifth principle reminds us that our history is important—it tells us who we are and why we do what we do. To understand why today’s engineers are as they are, we must understand better the history of theory and practice together—that is, engineering. As John Henrik Clarke, the American historian, said ‘History is not everything, but it is a starting point. History is a clock that people use to tell their political and cultural time of day. It is a compass they use to find themselves on the map of human geography. It tells them where they are but, more importantly, what they must be’. It reminds us that engineering education has placed little or no emphasis on the history of its disciplines—with the consequence that many of today’s engineers see little value in their heritage. An acronym reminder of these suggestions is PUPIL (Part, Unintended, Preparedness, Ingenuity and Learning). More fully stated ‘we are Part of a world of Unintended consequences for which we need to be Prepared through Ingenuity and Learning’. The first principle of Part—we are of this world not just in this world—is partially embedded in a better appreciation of the effects of moving away from the notion of the universe as a clockwork machine. This begs the first question that practitioners have to address in order to ‘ground’ their understanding before deciding how to act responsibly in any way that can harm others. ‘What do we actually know about our world?’

30

3 Grounding

The Clockwork Universe—A Discredited Myth? Not so long ago, even as late as the nineteenth century, scientists thought that the universe works like clockwork—a machine with reluctant secrets that slowly we are uncovering.4 The idea came out of the advances in science in the period after the Renaissance. Sir Isaac Newton’s laws implied that once the majestic clock had been set in motion future development was, in principle, entirely predictable. Since the laws were discovered through the munificence of God, then we had His permission to do whatever we liked to the world for our own use and convenience. Now most scientists would agree that the clockwork universe is a discredited myth—regardless of religious beliefs. But despite this rejection some of its effects are deeply embedded in our culture and manifest in diverse ways. For example, we still refer to the ‘laws’ of nature— implying that they are inviolable rules established by some external authority— mystical or otherwise. The word serves the illusion that the universe is orderly and bound by rules that we can discover and comprehend. Popularisers of science in the media refer to the fundamental laws of nature as though they are invariable when history tells us that they will change.4 The myth is implicit in the deterministic science that is still taught in our schools and universities. Determinism is the idea that all effects have a cause or more simply the future is determined by the past. Although everyday life and practical experience told a different story of vagaries and uncertainties, in 1828 Thomas Tredgold articulated the myth (unintentionally I am sure) when he defined civil engineering as ‘…being the art of directing the great sources of power in nature for the use and convenience of man…’ (my italics). The myth is still embedded in the way many scientists think about uncertainty as arising only through the randomness that stems from our ignorance of all of the necessary physical information.5 The adage ‘Save the Planet’—one of the most misleading and anthropomorphic statements ever made—embodies the myth. The planet is not just about us and we are not the planet. We are not in total control of the planet. If the climate does change as forecast, then the planet will be fine, but we won’t. What is under threat is not the existence of the planet—rather the danger is to the future well-being of all living creatures on the planet. Collectively, we have made life better for ourselves in so many ways but now in the twenty-first century we are only slowly coming to terms with the implications of rejecting the myth of the clockwork universe. We are discovering that we actually know with less certainty what we thought we knew. If you need proof then just think of all the events that have not turned out as we expected, such as the banking crisis of 2008, the spread of new types of terrorism, the rise of political populism in 2016, as well as flooding and other forms of damage from natural disasters, earthquakes and volcanoes. There are lots of smaller and less well-known examples that result in engineering failures. We all know that everyday life is unpredictable and the unexpected happens. We are living through perhaps the greatest ‘game-changing’ revolution in human achievement for a very long time. Quantum mechanics, complexity theory6 and

The Clockwork Universe—A Discredited Myth?

31

systems biology and engineering are altering the way we think and therefore what we can do. Scientists still refer to the laws of nature but only because they sense there is some order in our understanding of the physical world. But the rules we distill are complex and violable—they do depend on context and situation and even some things we think we understand turn out to be essentially unpredictable. The science of quantum mechanics now underpins all understanding of the physical world, but we can meet most of our practical needs without it—though we would not have laser scanners, light-emitting diodes, solar cells, electron microscopy, and lots of other modern equipment. Scientists talk in terms of a ‘theory of everything’—a framework of ideas that unite general relativity theory and quantum field theory— but the theory does not tell us what will happen tomorrow at the level of human existence. Nevertheless, quantum mechanics will probably continue to underpin new developments in quantum computing and nanotechnology and a whole host of yet unforeseen technologies. Most scientists now see a world of complex layers from the minute sub-atomic particle to the gigantic cosmos. And we humans are in the mix—we are of this world and our futures are bound up with it—as per the principle of Part. At each layer, there are interacting processes from which the properties at the next layer up emerge. In the new theory of complexity, something is complex if it is difficult to describe and predict. A complex thing cannot be fully understood by one person, so it is important that people collaborate in order to manage its performance. Many now distinguish between the words complicated and complex. A complicated thing has many interconnections, but the individual processes and connections are well defined and well understood—but perhaps only by a specialist. For example, if your complicated wristwatch develops a fault, you will need to go to a watch repairer. Aeroplane jet engines are very complicated but well understood by the engineers who design, build and maintain them. But they are so very complicated that they border on the complex. A complex thing consists of so many inter-related parts that the whole is more than the sum of its parts. Our bodies are complex things. Our cells are complex things. Engineered objects were once simple things (plough, bicycles, cathedrals and even cars) but modern things, like the Internet, are not simple—they are most emphatically complex. In a complex system, our actions may have the good consequences we intend. But there is also a distinct possibility of bad unintended consequences such as bullying and abuse on social media. When trains crash, bridges collapse and earthquakes bring down buildings, the law of the unintended often reveals deficiencies in human decision making as unintended consequences are exposed. Complex systems present us with new challenges. These new challenges cannot be tackled by any single group—they require a new kind of collaboration and tolerance that engages everyone—technophiles and technophobes, natural and social scientists, lawyers and economists, people and politicians. Whether we are aiming to improve our cities or deal with terrorism, we have to start by recognising that the future is uncertain. As the physicist, Niels Bohr is widely quoted ‘Prediction is very difficult, especially about the future’. Experts and non-experts alike are on a learning journey— an eventful odyssey into the future together. And we should never forget that an expert

32

3 Grounding

(doctor, lawyer, politician, engineer, scientist and economist) is only an expert in a narrow field—in everything else the expert is a non-expert just like the rest of us. If we cannot rely totally on our predictions without the distinct possibility of significant and harmful unintended consequences, then the principles of preparedness, ingenuity and learning will help. Historically, events have taken us by surprise, and we have often been slow to respond. Speed and agility of response depend crucially on our ability to collaborate. Yet we recognise that critical discussion, disagreement and dissent are a necessary part of progress. If we are to be free to disagree but also flexible enough to collaborate, then goodwill and good leadership to create a collective vision is crucial. Just as a top sports team needs a top-class coach so we need leaders with vision to articulate common goals and central values such as tolerance and good humour. Leaders can inspire us and release our innate resilience which depends on our ingenuity to spark and create ideas, to think through ways of responding to surprises and to learn through collaboration.5 If we are properly prepared, we can take advantage of unexpected good opportunities and protect ourselves by being robust, and sufficiently ‘fleet of foot’, to adapt and navigate the bad.

A Learning Odyssey The present is just a brief moment in moving from past to future. Everyone, born and the unborn, has a vested interest in our learning odyssey—‘learning to learn’ how to collaborate our way through the future. Perhaps the only thing we can be sure of is that we cannot be sure. But we can agree on some common goals—at the top of which is survival. In War and Peace Leo Tolstoy wrote ‘We can know only that we know nothing. And that is the highest degree of human wisdom’. For sure we need that wisdom. There is no assurance that the future will be a linear extension of the past. Venturing into an uncertain future, however, does require us to understand the past and how it became the present so that we can build a platform to think about how we adapt in the future. The learning odyssey of each chapter of this book will address the engineering of some of our most basic needs and wants such as shelter, moving around, communicating, defending ourselves and well-being. Then, in the final chapter we synthesise the lessons from the past and present into some of our important and not inconsiderable future challenges. As we look at the past, we find that there was the game-changing period around the Renaissance—a fuzzy dividing line between activity before and after—roughly from the fourteenth to seventeenth centuries. Before the divide, the roles of what we now understand as engineer, scientist, craftsman, artist and architect were not clearly distinguished—they could be taken by one man usually at different times for different works. The word engineer (or Medieval Latin ingeniator) tended to be used about people who were concerned with weapons of warfare—but not entirely. For example, in England, Ailnolth was the first to be called ingeniator but was also overseer of royal buildings for Henry II for over 30 years from 1157 and

A Learning Odyssey

33

Leonardo da Vinci was Ingenarius Ducalis (the Duke Master of Ingenious devices) for the Duke of Milan around 1482. After the Renaissance, the influence of the new scientific thinking grew. People began to develop and use the deterministic science created by Newton. As the new knowledge developed, people started to specialise. For example, architecture began to be separated from engineering. Architecture became the design of concept and hence associated with the creative arts. The idea of ‘fine art’ gained ground. Italian artist-engineers such as Michelangelo who were, during their early careers, regarded as artisans became revered as inspired and gifted fine artists. Today, very few of us could afford to buy a Michelangelo painting and hardly anyone would remember him as an engineer or even an architect yet, for a time, he was responsible for St. Peter’s Basilica in Rome. Engineering began to be associated with the craftwork of making and repairing. Engineers were seen to be merely executing what the talented designers instructed them to do—work that was mundane and not creative—manual labour or craft skill. Later it became the simple application of the known burgeoning science to convert the vision of the designers into reality. There were notable exceptions, however—as there are to any rule. The steam engine was developed in the seventeenth century by creative practitioners without any science—as had the stone arch by the Romans centuries before. The science of thermodynamics developed out of the need to understand steam engines but now of course is invaluable in designing new engines—just as the science of structural analysis now helps us design new arch bridges. Since the new science was understood as the uncovering of universal truth— revealing how the clockwork machine works—then discovering scientific laws was clever but not particularly creative. It was the ‘laying bare’ of nature rather than the result of individual inspiration. Applying the laws was straightforward because they were inviolable. We simply had to use them to get what we want—requiring some intellectual effort but without creativity. The success that followed exacerbated the view. Steam engines, railways, the internal combustion engine, electricity, telegraph, radio, cars and aeroplanes television and jet engines all reinforced the idea that we could do whatever we wanted. The engine driver of the steam locomotive engine entered public consciousness as the engineer—the words were too alike to be otherwise. Science took over and became king. Specialisation through science has led to three important and unfortunate developments. First, the engineering professions have fragmented with a consequent loss of connection between the ‘big picture’ and the detail in both engineering education and practice. Second, the level of exchange of information and experience between disciplines has reduced significantly because each has developed their own ‘language’ or technical jargon. Third, and perhaps most important of all, the dominance of science has resulted in a loss in the valuing and nurturing of ingenuity through practical wisdom.

34

3 Grounding

Problem Solving with Practical Wisdom But it hadn’t always been like that. In 330 BCE Aristotle wrote in his Nicomachean Ethics: Some people who do not possess theoretical knowledge are more effective in action (especially if they are experienced) than others who do possess it.

I wonder if Aristotle’s observation rings true to you today. I feel that we undervalue practical experience and overvalue academic ability. The academic/vocational distinction mars our educational systems. It is commonplace nowadays to hear people say, ‘Tommy isn’t academic—he’s much more practical’ or ‘Gianna is bright but no good with her hands’—as though being practical and academic are mutually exclusive. They certainly aren’t exclusive in your local dentist or hospital surgeon and neither are they in engineers. However, the notion of practical wisdom as articulated by Aristotle been lost in the mists of history. Consequently, we just do not recognise, value or nurture talent or genius (Latin ingenium) in practical matters that involve science even though essential to making things work. We recognise individual talent where science has little impact such as sport (the soccer skill of Ronaldo), entertainment (the vocal quality of Elton John), arts (the brushstrokes of Michelangelo) and crafts (the pottery of Grayson Perry) but not in technical education or trades (although in the nineteenth-century people such as I. K. Brunel were indeed famous). Now practical people are often made to feel inferior compared to those with more theoretical knowledge—especially in schools and universities. The practical wisdom associated with practical problem solving has been made the handmaiden of science—through the influence of the myth of the clockwork universe. The myth continues to reinforce, in some people, ‘technology triumphalism’— the idea that modern digital technology will remedy all societal issues. It is perhaps a manifestation of the more general western triumphalism that dominates modern ‘globalisation’ and against which the trend for ‘populism’ is a reaction. Engineers are natural optimists—they have to be to tackle some of the daunting projects like sending men to the moon, building Crossrail, and tunnelling under the English Channel. None of these is tasks for the faint-hearted. But they have to guard against triumphalism. Optimism has to be tinged with realism—‘techie’ triumphalism is a risk. For example, Egyptian Wael Ghonim, who helped usher in the Arab Spring, is reported to have said ‘If you want to liberate a society, just give them the Internet’. Eric Schmidt, the executive chairman of Google, and his co-author, Jared Cohen, write7 : ‘The best thing anyone can do to improve the quality of life around the world is to drive connectivity and technological opportunity’. The title of the 2005 BBC Reith Lectures by Alec Broers at that time the President of the Royal Academy of Engineers was ‘The triumph of technology’. Although he did emphasise the need to avoid complacency, he said that technology can solve our problems if we scientists and engineers engage with the public to explain why we are right. This kind of thinking will let us down because the impression that scientists and engineers have all of the answers is seriously wrong. In truth, they have many

Problem Solving with Practical Wisdom

35

answers that are good but only in the contexts they understand. We all have a stake and a contribution to make and a duty to collaborate to find solutions. Unintended consequences (like climate change and pollution) could prevail if we are not very careful. I think that implies that we all need to think much more critically about what we know and the way that influences how we tackle our collective challenges, manage our practical problems, react resiliently to surprises, and above all address our individual and collective purposes together. It requires transdisciplinarity as the integration of diversity across all divisions by active collaboration. We can start this process by examining what is common and at the core of solving all practical problems. We each have our own ways—some good and some not. Typically, analysts will identify four to seven stages we ought to go through. But because we often don’t consider them formally or even informally, they may well go unnoticed in everyday decisions. However, in formal professional decision making where decisions may be called into question at a later date, then they must be worked through systematically. Of course, the stages are not neatly, discretely and tidily defined—there is much toing-and-froing, hesitancy, blind alleys and dead ends. Nevertheless, we can outline, say, a simple four-stage process as the plain common sense: plan, do, check and act. Planning is thinking about where you are and where you want to be. Doing is organising to implement your plan. Checking is assessing how well you are doing and investigating when things go wrong, learning and revisiting your plan. Acting is starting over in a new situation. A typical seven-stage process is just a little more detailed: identifying the problem, framing to set some goals, conjecturing to explore possible solutions, thinking to select the best one, sharing ideas to find ways to test the choice (e.g. think through the possible unintended consequences), communicating to implement (act) and then diagnosing what might be wrong and being prepared to review the consequences (intended and unintended). Then we get a new state of affairs, a new problem which we reframe and round we go again. In (Fig. 3.1a), these stages are shown as a loop—a problem solving loop that we go round and round in a spiral through time (Fig. 3.1c). For example, you are hungry. A snack will suffice. You could eat a sandwich or a wrap. You select a cheese sandwich. You check to make sure it’s ok to eat. You eat but then realise you are still hungry. So, round the spiral you go again perhaps this time choosing a chocolate bar. There are many variations on these four to seven stages. One particularly important one in the history of science was a six-stage scheme proposed by the philosopher Karl Popper in 1935. He wanted to capture the evolutionary growth of scientific knowledge8 but his analysis can also be applied to engineering. His six stages are identifying the problem, conjecturing a solution, deducing a testable proposition from the conjecture, devising a test of that proposition, carrying out the test and, depending on the result, being prepared either accept the conjecture for the time being or reject as being false. Then we have a new problem and round we go again. Applied to engineering, the problem would be to address a need (such as a new building), conjecture a particular design solution, devise ways to test that solution theoretically and practically, carry out the tests (analyse or test the solution) and either adjust or adopt the design solution.

36

3 Grounding

Fig. 3.1 Decision loops

Many scientists and engineers reject that such a neat scheme captures what they do—nevertheless, there is a strong logic behind Popper’s scheme which many find helpful. Two other points are worth noting—the second one is particularly controversial amongst philosophers and scientists. First Popper’s problem is the innate curiosity of the scientist—our need to understand, our need to ask interesting questions. The engineer’s problem is different— finding a way of satisfying a practical need or want. Popper wanted to identify how understanding develops. He recognised evolutionary growth and he saw ingenious problem solving at the heart of that growth. To him, a conjectural solution was a scientific theory or hypothesis concerning a question. He said that scientists deduce a proposition from that conjecture that they know they can test by an experiment. For example, imagine we want to explain the trajectory of a ball that we throw high in the air. We conjecture that Newton’s Laws apply. We use the laws to predict some important characteristic of the trajectory such as the length of the throw under some specific conditions, such as no wind. Then we actually throw a ball in a way that matches those conditions and measure the distance we throw. We compare the measurements with our predictions and accept or reject the conjecture. Over the years, many people have done this and got good agreement. So, we can use the theory with some confidence, for example, to design missiles for warfare—though in important applications like that we have to do lots more tests. Note that Popper’s scheme says almost nothing about how Newton conceived his laws though he clearly was, like all of us, building on what has gone before. For example, his first law was previously formulated by Galileo and Descartes though in a slightly different form (for example, Descartes attributed force as something caused by God whilst Galileo clearly grasped the idea of force by connecting force and motion).

Problem Solving with Practical Wisdom

37

The second point of note with Popper’s scheme is trickier. How do you decide after the test that the result agrees with (confirms) or doesn’t agree with (refutes) the proposition and hence the conjecture? Popper argued quite rightly from a logical point of view that one refutation proves that a conjecture is false. In a clear-cut strict logical sense, he was right if we can be precise in our statements so one black swan disproves the hypothesis that all swans are white. But in scientific experiments results are often approximate so that if we do a number of tests, we may get slightly different answers. Consequently, scientists often find difficulty in agreeing what actually constitutes a refutation without some kind of tolerance limit. Even if we get a series of confirmations, we cannot assert that the conjecture (theory) is absolutely true in all situations and all contexts because we can never be sure if there is some test we haven’t thought of before that will refute the conjecture. For example, Newton’s laws do not explain the precession of the planet Mercury—a refutation that eventually led to Einstein’s theory of Relativity.9 So, belief in the absolute truth of any theory has to be an act of faith, i.e. one not based on incontrovertible proof. There is no shortage of suggestions for problem solving loops that represent how products are engineered. The simplest is perhaps design, make/build and operate with maintain as a further stage often tagged to the end. An eleven-stage suggestion is the cycle shown in (Fig. 3.1b). Through it and successively through ingenuity new practical solutions evolve just as new theories do. All of these suggestions for problem solving loops seem, on the face of it, to be quite different. In detail, they differ but in essence they can be boiled down to variants on five core stages: problem, possible solutions, select one, test and use, new state of affairs and new problem. Of course, engineers in different companies and industries build much more elaborate schemes from this base. But by identifying these five core stages, we can see that many of the activities we think of as being very different do indeed have a common nucleus. My five principles apply directly to the stages. You will recall that ‘we are Part of a world of Unintended consequences for which we need to be Prepared through Ingenuity and Learning’. The principle of Part helps us to look for layers of issues in our problem in order to cope with and manage complexity. The principle of the Unintended reminds us to look out for the surprises and be aware of possible unknown unknowns when we are creating possible solutions. The principle of Preparedness tells us to evaluate and select solutions that include a capability to cope with surprises, and constantly ‘reading’, diagnosing or reviewing the state of affairs. The principle of Ingenuity prompts us to think innovatively, resourcefully and inventively but particularly when we are testing our possible solutions to look for flaws. Finally, the principle of Learning emphasises consciously to learn from what we do as we examine the new state of affairs and redefine our problem ready for the next loop. I have said that engineering as ingenious problem solving is a central part of what it is to be human—but as individuals we engineer informally, and we may not consciously go through these problem solving loops. Professional engineering problem solving processes differ in three important ways.

38

3 Grounding

First, engineers have to be formal in the sense that all decisions have to be publicly justifiable. They, like all professionals, have a legal duty of care—an obligation under the law of tort to a reasonable standard of behaviour when acting in a way that could harm others. The test of reasonableness would ultimately be decided in a court of law commensurate with the expertise claimed by the engineer. It is a determinant of negligence where a professional has represented him or herself as having more than average skills and abilities. Second, the reason for the formality is that the outcomes of engineering work are often of massive scope and influence—just think of transportation networks and big bridges, computers and the Internet—the work of engineers has a huge effect on everyone. Third, professional engineering problem solving is about configuring flows of energy to achieve desired outcomes and purposes—though that is not the way the process is normally articulated. Energy is the key integrating concept that opens up the commonality between the various arms of a fragmented profession. Later, we will delve into the problem solving at the heart of how energy flows are used in the medical engineering of health and well-being (including some details on how Ben’s pacemaker came into being), buildings and bridges, transport, and warfare. But we start with that most ancient form of engineering—providing shelter. End Notes 1. The UN Secretary-General António Guterres continued ‘The world reached several dire milestones in 2017. The economic costs of climate-related disasters hit a record: $320 billion. Energy-related carbon dioxide emissions rose 1.4 per cent, to 32.5 gigatonnes—a historic high…. And I am beginning to wonder how many more alarm bells must go off before the world rises to the challenge. See https://www.un.org/sg/en/content/sg/press-encounter/2018-0329/secretary-generals-press-encounter-climate-change-qa (Last accessed February 2019). 2. What Aristotle meant by phronesis is open to interpretation. I see it as a means towards an end arrived at through moral virtue. It is the capacity for determining what is good for both the individual and the community. It is a virtue and a competence, an ability to deliberate rightly about what is good in general, about discerning and judging what is true and right but it excludes specific competences (like deliberating about how to build a bridge or how to make a person healthy). It is purposeful, contextual but not rule-following. It is not routine or even well-trained behaviour, but rather intentional conduct based on tacit knowledge and experience, using longer time horizons than usual, and considering more aspects, more ways of knowing, more viewpoints, coupled with an ability to generalise beyond narrow subject areas. See more at https://blog.oup.com/ 2014/07/practical-wisdom-vsi/ (Last accessed February 2019). 3. In his book ‘The Ingenuity Gap’ Thomas Homer-Dixon defines ingenuity as ‘instructions that tell us how to arrange the constituent parts of our social and physical worlds in ways that help us achieve our goals’. See

Problem Solving with Practical Wisdom

4.

5.

6.

7. 8. 9.

39

https://homerdixon.com/wp-content/uploads/2017/05/Homer-Dixon-TheIngenuity-Gap-1995.pdf (Last accessed February 2019). Snobelen, SD. (2012) The Myth of the clockwork universe, Chapter 6 of The persistence of the sacred in modern thought, ed. Chris L. Firestone and Nathan Jacobs. University of Notre Dame Press, 2012, pp. 149– 84. See https://isaacnewtonstheology.files.wordpress.com/2013/06/the-myth-ofthe-clockwork-universe.pdf (Last accessed February 2019). See http://myengineeringsystems.co.uk/5-axioms/ (Last accessed February 2019) and also Blockley, DI, Godfrey, PS. (2017), Doing it Differently: Systems for rethinking infrastructure, 2nd edition. ICE Publishing, London. For a good introduction to quantum mechanics I recommend Penrose, R. (Revised Edition) (2016), The Emperor’s New Mind, Oxford University Press For complexity theory I recommend Cohen, J., Stewart, I. The Collapse of Chaos, Penguin Books, London For systems biology I recommend Noble, D. (2008) The Music of Life, Oxford University Press For engineering I recommend Blockley, DI, Godfrey, PS. (2017) Doing it Differently (2nd Ed) ICE Publishing, London. Schmidt, E., Cohen, J. (2014) The New Digital Age, John Murray. Popper, KR. (2nd Edition 2002) Conjectures and Refutations, Routledge, London. The planet Mercury does not move as predicted by Newton’s Laws. Mercury’s orbit around is approximately an ellipse but with a changing closest approach to the sun—the orbit of which is called a precession.

Chapter 4

Dwelling

Unintended and not amended, inadvertent will be lamented

Ronan Point and Grenfell Towers Ivy Hodge fancied a cup of tea very early one morning in May 1968. She got out of bed and sleepily made her way to the kitchen of her flat, on a corner of the 18th floor of a 22-storey apartment block in Newham, in London’s East End. The building, Ronan Point, was only two months old. Ivy lit a match. In the massive gas explosion that followed, the load-bearing walls of her flat were blown out. Fortunately, she survived, though she suffered burns as she was blown across the room. But the walls of her flat supported the rooms above and so within seconds the entire corner of the building collapsed—like dominoes in sequence—a phenomenon known as progressive collapse. The explosion and collapse killed four people and injured 17 of the other 240 residents. Eighty families fled their homes, many in their nightclothes. A young mother was left stranded on a narrow ledge when the rest of her living room disappeared. In 2017 the Grenfell Tower fire in London killed 71 people. The fire started by accident, due to an electrical fault in a flat on the fourth floor, but then escalated more rapidly than fire experts expected. From the early pictures, the exterior cladding was suspected as being inadequate. The fire spread progressively through the cladding. Fortunately, the building did not collapse—had it done so the death toll would have been much higher. At the time of writing the Grenfell Tower Inquiry is still hearing evidence and is expected to report in 2020. A review of the Building Regulations for the UK is underway as is a criminal investigation. An interim report1 concluded that the ‘clarity of roles and responsibilities is poor. Even where there are requirements for key activities to take place across design, construction and maintenance, it is © Springer Nature Switzerland AG 2020 D. Blockley, Creativity, Problem Solving, and Aesthetics in Engineering, https://doi.org/10.1007/978-3-030-38257-5_4

41

42

4 Dwelling

not always clear who has responsibility for making it happen’. Although the trigger event for the Grenfell Tower disaster was quite different to that at Ronan Point the incubation of the unintended consequences seem, at the time of writing, to be similar and the effect is clearly the same—loss of lives and homes. Unsurprisingly the incidents at Ronan Point and at Grenfell Tower led to a loss of public confidence in high-rise residential buildings in the UK. After Ronan Point the UK Home Secretary at the time James Callaghan immediately ordered an inquiry. He chose a tribunal led by Hugh Griffith QC with Professor Sir Alfred Pugsley, a civil and aeronautical engineer (and one of my predecessors as head of civil engineering at the University of Bristol, UK), and Professor Sir Owen Saunders, a mathematician and mechanical engineer and a former Vice Chancellor of the University of London. The Tribunal decided that the problem was inherent in the design of the building. Large reinforced concrete panels were arranged to form boxes stacked upon each other like a house of cards. The structural integrity of the building relied on good joints between the panels—but the inquiry found that they were deficient; the panels were not tied together sufficiently well. In other words, if any one of the panels was removed for any reason, just as if you remove a card from a house of cards, there is no way that the building can stand. In more technical language there is no alternative way for the forces and loads in the building to flow to the ground and so, just like a house of cards, the entire structure collapsed progressively in sequence. The Tribunal also realised there was even more to learn and do. Although the immediate trigger was a gas explosion, anything that might remove a wall panel needed to be thought through. They deduced that the pressure on the walls would have been quite low because Ivy’s injuries were thankfully not as severe as they might have been. They calculated that the strength of the panels wasn’t even adequate for the expected wind loading, or for a fire that might cause a panel to bow out of alignment. The Tribunal therefore recommended that all existing blocks in large panel construction over 6-storeys in height should be appraised and strengthened as needed. They said that the gas supplies should be cut off in any blocks judged susceptible to progressive collapse. They urged that the building regulations and the relevant codes of practice be revised. However, they also said, once the risk of progressive collapse was removed, there was no reason to prohibit the use of gas in high buildings and no reason why large precast concrete panels should not be used. Ronan Point was partly rebuilt with strengthened joints. But it became clear that there were even more defects. Gaps were found between floors and walls severing the fire breaks that made the fire and acoustic ‘compartments’ necessary to stop a fire from spreading. Even more structural problems were uncovered—for example, the forces passing from one wall panel to the next were not spread evenly and instead were transmitted by two steel rods. Some strengthening brackets were not properly attached. Eventually, the concerns became so great that the council evacuated the building. They had it demolished in 1986 but as they did, even more defects were found. All through the project, corners had been cut and responsibilities were unclear. Every one of us recognises our important and fundamental need for somewhere to live and work. Some place that gives a sense of belonging and community. The questions for this chapter include ‘How have modern buildings evolved from the first

Ronan Point and Grenfell Towers

43

caves and primitive huts?’ What are the roles of architects and engineers? Have those two professions diverged—even when individual professionals and companies they work for, still must work closely together? Can someone with no technical training, learn to look at a building or other structure and ‘read’ how it ‘stands up’ or works as a structure as well as appreciate its external aesthetic and architectural appearance? What are some of the technical difficulties that engineers face, such as wind and vibration? Can we continue to expect our buildings to be ‘as safe as houses’ when— as the stories of Ronan Point and Grenfell Towers illustrate—sometimes they fail? How do we avoid the type of progressive collapse that happened at Ronan Point? Accidents and disasters are obvious examples of a failure to apply the principle of the Unintended. These events are not supposed to happen so understandably many jump to the conclusion that someone must be at fault—a natural reaction of victims is to blame. The principle of Preparedness is about being ready to respond to events of this kind, to minimise impacts and to get back to something like normal life as soon as practical—even though ideas about what is normal may change. Part of the response is to look for evidence of negligence and, of course, if found then those responsible must be brought to account. But often the underlying chains of events are subtle, enigmatic and systemic (i.e. a characteristic of the whole) with individuals acting in good faith. The reasons for an accident emerge in layers from complex interactions between events at lower layers—according to the principle of Part. The principle of Learning is about winkling out those events and their emergent properties. The sociologist Barry Turner spent his life studying them. Barry’s untimely death at 57 came when he was at the peak of his powers working across traditional academic boundaries. He started life as a design engineer in the chemical industry but began his academic career as a mature student. He went on to research the theory and practice of organisations during which time he and I worked together with the psychologist Nick Pidgeon. In his book ‘Man Made Disasters’2 [with Nick helping on 2nd edition], Barry concluded that accidents don’t just happen they ‘incubate’. By that he meant that a trickle of events and smaller decisions, each with unintended consequences, build, interact, emerge and accumulate—some of them from punching through the as-is contextual assumptions mentioned in Chap. 2. It’s rather like blowing air into a balloon. Each puff of air represents an unintended consequence. The growing balloon mirrors the serious situation—the pressure of the air is the emergent measure of an accident-waiting-to-happen. If nothing is done to relieve the situation (by letting some air out of the balloon—managing the effects of the unintended consequences) then a ‘trigger’ event (a pin in a balloon—a gas explosion or fire) will cause an accident (burst the balloon—bring down a building) sometimes with consequences much larger than the trigger might suggest. Ronan Point was a tragic event—very few apartment blocks collapse. With hindsight, the Barry Turner incubation period was there to see—a series of unfortunate circumstances leading to an ‘accident-waiting-to-happen’ and a trigger event. The incubation process started when the UK Government Housing Subsidies Act of 1956 introduced subsidies to local councils to build tower blocks over five storeys. The shortage of skilled labour meant that choosing a ‘so-called’ industrial building technique was natural—quicker and needing a smaller labour force. The method was

44

4 Dwelling

innovative, inventive and ‘cutting-edge’ with the promise of considerable savings by prefabricating large concrete panels off-site, using them both for the walls and floors and joining them together securely on site. The idea of progressive collapse was not on anyone’s radar. There had previously been some unexpected and sudden collapses of ancient stone towers but no widely known cases in domestic tower blocks. For example, the Campanile in San Marco Square in Venice collapsed suddenly in 1902 (the present tower is a reconstruction) but there was no apparent connection between that event and the proposal for Ronan Point. The UK Building Regulations contained no mention of progressive collapse. One departure was that Ronan Point was to be taller than any previous building using this technique—but that didn’t seem to be an issue at the time. No potential for unintended consequences was foreseen and there was no sense that particular care was needed to test innovation. The contract was led by a building company who employed structural engineers, so the professionals were not directly accountable to the client—again, on its own, not an issue but, in the event, a contributory factor. The checking of the calculations by the design team and by the local council was limited—that was poor practice. The young site engineer supervising the construction was inexperienced and not fully qualified as a professional structural engineer. There was no criticism of him but putting him in charge without enough experience was poor practice by his employers. Whilst the construction process, seemed at the time, to be generally good there were defects in the joints between the concrete panels—again evidence of slipshod practice. During the investigation, one of the nuts used to fit Ivy Hodge’s gas cooker was found to be flawed, although it had worked satisfactorily for some time, and was judged to be a random fault. Rather surprisingly no one had smelled a gas leak. When all of these factors are put together the building became an ‘accident-waiting-to-happen’. Ivy lighting her match was the trigger. Throughout the incubation period Unintended consequences were rife in all of the vertical layers of the problem, as per the principle of Part—from the UK housing shortage after WWII, in the top layer to the technical detail of the fixing of the joints in the bottom layer. The principles of Preparedness, Ingenuity and Learning were missing because no one thought they were needed—decisions were made without realising that innovation demands a need for vigilance and good practice. After the enquiry, report was published significant changes were made to the regulations that cover buildings in the UK. Major changes were made to the Code of Practice for wind loading because high winds could have removed one of the panels. Of more immediate relevance for the residents of Newham, was that nine blocks and many other buildings were demolished. Unfortunately, through media reporting, many people were made to worry unnecessarily about all forms of highrise structure when there was no need—because the structures of their buildings were quite different. Ronan Point was a game changer for structural engineers. The lessons about structural design and construction have been learned but the Grenfell Tower tragedy has exposed possible similar engineering safety concerns. Not this time regarding the structure itself as at Ronan Point, but rather relating to those responsible for fire safety, the use of a form of cladding, co-ordinating responsibilities and checking that

Ronan Point and Grenfell Towers

45

buildings are built as designed. Grenfell Towers may turn out to be another defining moment. People living in high-rise blocks have a right to feel safe. It is natural to feel anxious if you live in a tower block like Grenfell—even if the form of cladding is quite different. The people responsible for the safety of those blocks need to demonstrate that the residents are not at undue risk. Assurances will carry little weight until a full review of the Building Regulations and their use is complete and consequent changes made. The Grenfell Tower fire was a tragic surprise to some professionals but to others, it was another ‘accident-waiting-to-happen’. The underlying story behind Barry Turner’s incubating period will hopefully emerge and will be layered with politics at the top (such as the effects of Government austerity policies) and technical issues underneath (such as changes in the design of cladding against fire becoming out of alignment with the proper allocation of responsibilities and accountabilities). Whatever the underlying causes the tragedy is a deadly reminder that politicians, civil servants and building professionals have to be vigilant and work together in managing potential risks and robust in ensuring that vulnerabilities are not inadvertently introduced. People with responsibility must be competent and must ensure people who are accountable to them are also competent. One suggestion that may emerge is that the entire works should be signed off by a ‘guiding mind’.3 This would be a properly qualified professional whose job would be to ensure that all aspects of a project are integrated and, most importantly of all, that the highest priority is achieved—the safety of people.

Shelter Without shelter we are vulnerable to the exigencies of our natural world. We need somewhere to feel safe, to make our homes and places to live and work. Imagine living as ice age humans in caves and in tents made from animal skins and bones—difficult for most of us to visualise with our modern comforts, though not so difficult for ‘rough sleepers’ and people living in desperate poverty and hunger in too many parts of the world. Only slightly better would be to live in one of the first towns we know about—the mud-brick houses in Çatalhöyük (in modern Turkey) around 7,500 BCE. Farming helped—people began to make huts made of stone or wattle and daub and they built thatched roofs. Bronze Age (3,500–1,250 BCE) people lived in round wooden huts. Knossos is probably Europe’s oldest city—at its peak in about 1,700 BCE there may been 100,000 people living there. The Celts in France and Britain, around 650 BCE, lived in round houses with wattle and daub walls and thatched roofs. The historic buildings of the Greeks and Romans such as the Acropolis in Athens, and the Pantheon (Fig. 4.1) and Colosseum in Rome are justly famous. The building of temples and monumental structures developed from the massive columns and lintels of the Greek period to the arches, vaults and domes of the Romans. The roof of the Pantheon, built in 120 AD, is an amazing 44 m diameter spherical concrete

46

4 Dwelling

Fig. 4.1 The Pantheon, Rome. Image by Richjheath, Public domain via Wikipedia Commons

dome. The thrust on the walls is reduced by using lightweight aggregate for the concrete with smaller thickness towards the crown. It is a highly intelligent example of integrated architecture and engineering. Domes are effectively three-dimensional arches. Arches were a significant advance because they rely on a supporting structure (the falsework or scaffold) to hold up all the stones or bricks until the full arch is in place. Once in place they will stand for a very long time if the foundations don’t move. The reason is that an arch is stable because of its shape rather than being made of strong materials. The self-weight together with the weight of anyone crossing the arch is carried by a compressive force in a line down through the arch to the foundations. That line of force is known as a ‘thrust line’ and as long as it lies within the shape of the arch then the arch will stand. The Romans used the arch to great effect—perhaps the most impressive, apart from the Pantheon, was the Pont du Gard (Fig. 8.2) aqueduct in France. The Romans also made concrete using naturally occurring volcanic pozzolana ash mixed with quicklime, aggregate and water. The feats of Roman engineering were the province of a tiny elite. Rich Romans lived in a domus—very grand, with marble pillars, statues, plaster or mosaic walls and floors. But the majority poor Romans lived in insulae of apartment blocks, grouped around a central courtyard. They were made of wood and mud brick with no heating or running water and were dirty, noisy and unhealthy places to live. Then the whole Roman way of life was destroyed after the fifth century and there was a general decline across Western Europe at all social levels. In Renaissance Italy, patrons began to want monumental buildings. They slowly recognised that artisans, who had always worked under the direction of guilds or

Shelter

47

the church, were not only skilled technicians but also creative thinkers, discoverers and inventors. The expanding numbers of bourgeoisie began to acquire their works as status symbols and the idea of ‘fine art’ gained ground. Leon Battista Alberti, Donato Bramante, Filippo Brunelleschi, Leonardo da Vinci, Raphael and Michelangelo were polymaths working as artists, poets, philosophers, architects and engineers. Brunelleschi was responsible for the dome of the Duomo (cathedral) in Florence in the fifteenth century, Bramante, Raphael and Michelangelo were all architects at various stages for St Peter’s in Rome. The Duke of Milan appointed Leonardo da Vinci as Ingenarius Ducalis (the Duke Master of Ingenious devices). As a result, respect for craft skills began to give way to conceiving, drawing and disegno (drawing and design). Andrea Palladio began to design buildings influenced by the Roman Vitruvius whose books of architecture were rediscovered in 1414 by Florentine Poggio Bracciolini. All of Palladio’s buildings are located in what was then the Venetian Republic. They are now the world heritage sites of the City of Vicenza and the Palladian Villas of the Veneto (Fig. 4.2). His four books of architecture gained him wide recognition. Inigo Jones brought the Palladian style to England in the seventeenth century with the Queen’s House at Greenwich and the Banqueting House, Whitehall, in London. Like many men of the period he was multi-talented, for example, he made major contributions to stage design by his work as theatrical designer for several dozen masques (a form of courtly entertainment with music and dancing), most by royal command and many in collaboration with Ben Jonson.

Fig. 4.2 Palladio’s Villa Rotonda, Veneto, Italy. Image by Stefan Bauer CC BY-SA-2.5 via Wikipedia Commons

48

4 Dwelling

Theory and Practice The turning point for the engineering of the structure of dwellings was the period after the Renaissance. New science brought new opportunities through specialisation but also fragmentation through division. Up until that time the only science for practical work was arithmetic and geometry. Rules for building were therefore often stated using proportions—for example, Palladio chose room sizes with proportions such as 3:4 or 2:3 with ceiling heights half the sum of the width and length. But progress in science during the Renaissance period and after was beginning to change that. For example, in the sixteenth century, Simon Stevin developed what was to become the triangle of forces and the decimal system—both are basic tools of modern engineering. After Stevin new science was to follow. In the seventeenth century, Galileo Galilei estimated the breaking strength of a timber beam and did some experiments. But it was Sir Isaac Newton who formulated the three laws of motion that became the fundamental theoretical tools of the modern engineer. The new science had only minor impact on engineering practice until the middle of the eighteenth century. One of the early innovators was Charles Coulomb, an engineer in the French army, who is perhaps better known for his discoveries in electricity and magnetism. Coulomb’s law that describes the forces between static and charged particles was published in 1784. Beside his electrical work, he laid the foundations for a theory of soil mechanics essential for the design of foundations of buildings and bridges as well as earthworks. Then in the nineteenth century, Frenchman ClaudeLouis Navier established structural analysis as a science and Italian Carlo Alberto Castigliano was one of the first to realise the importance of energy in structures and, for his doctorate, developed theorems later generalised by German Freidrich Engesser. Iron and steel were becoming available in commercial quantities—creating the potential for new types of structure. Abraham Darby I replaced wood with coke to smelt iron in substantial quantities from 1709 onwards. His grandson, Abraham Darby III, erected the famous cast iron arch at Ironbridge over the River Severn in 1777–9. In 1784 H. Cort produced wrought iron in a coal-fired flame ‘puddling’ furnace. A form of the puddling process was known to the Chinese Han dynasty around the first century A.D. Modern ‘puddling’ was one of the techniques developed to make bar or wrought iron from pig iron without using charcoal. It was called puddling because molten iron was blown by a strong current of air and stirred with long rods which were consumed in the process. Cort went on to invent grooved rollers for making bars and plates. Previously they had to be cut from hot strips and hammered into shape. John Smeaton was one of the first engineers to use cast iron for windmills, water wheels and pumps. His novel cast iron beams were I-shaped but had small top flanges and large bottom flanges. Flanges are the projecting rims on the top and bottom of the I-shape. Smeaton knew that they are important determinants of the strength of a beam in bending as they concentrate material away from the centre of cross section. In 1805 Armley Mill near Leeds was one of the first and largest woollen mills in the

Theory and Practice

49

world to be built using cast iron circular columns. Inverted T section cast iron beams supported a shallow brick arch floor, but these were very heavy. Cast iron columns with timber floors became standard for many mills and factories. An altogether more adventurous and iconic structure was built in 1851—the Crystal Palace (Fig. 4.3) in Hyde Park London for the Great Exhibition. It was effectively a massive greenhouse built in cast iron and glass prefabricated components designed by Sir Joseph Paxton with contractors Fox and Henderson. Paxton had pioneered building greenhouses at Chatsworth in Derbyshire and in 1836 had built a huge glasshouse there called the Great Conservatory. In 1856 Sir Henry Bessemer patented a way of replacing the puddling process for making steel by blowing a blast of air through the fluid pig iron. Consequently, steel became available in large quantities at economic prices. Unfortunately, some of the first steel was of poor quality but the problems were soon solved. By the early 1900s steel was commonly used for floors in building but usually supported by cast iron columns or brick walls. The date of the first steel framed building in the UK is uncertain because the words iron and steel were often used interchangeably. The National Liberal Club in London built in 1879 had columns and girders of steel and iron. In the USA the first tall building with a steel and iron frame was the Home Insurance Building in Chicago built in 1885. The first fully steel framed building in the UK was probably the Royal Insurance Building, Liverpool built between 1896 and 1903.

Fig. 4.3 Engraving of the front entrance of the Crystal Palace (Wikipedia Commons) London built for the Great Exhibition (Wikipedia Commons) of 1851. Public domain via Wikipedia Commons

50

4 Dwelling

Another newly emerging major structural material was reinforced concrete. The Romans knew that concrete is strong when compressed but weak when pulled in tension, but the use of cement declined after the fall of Rome. John Smeaton in 1756 experimented with hydraulic lime for the Eddystone Lighthouse but it wasn’t until 1824 that Joseph Aspdin produced a cement that he named after Portland stone. François Coignet had inherited a family business and opened up a new cement factory near Paris in 1852. In order to promote his venture, he did some experiments with ironreinforced concrete and went on to build a house made of béton armé or reinforced concrete. A few years later a French gardener Joseph Monier patented reinforced flowerpots. But there was, as yet, no understanding of how the steel rods or mesh helped to make the structures stronger. American New Yorker Thaddeus Hyatt took out a patent for iron-reinforced concrete floors in London in 1878. By 1903 confidence in the experimental tests was such that the first reinforced concrete skyscraper of 16 storeys was the Ingalls Building in Cincinnati, Ohio (Fig. 4.4). An international committee was set up in Germany in the early 1900s to report on experiments and Fig. 4.4 The Ingalls building in Cincinnati, USA. Image by Rdikeman CC BY-SA-3.0 via Wikipedia Commons

Theory and Practice

51

to develop the much-needed theory. Only in 1907 was the first textbook published by Frederick Turneaure and E.R. Maurer in USA setting out a history of previous experiments and theory.

Fragmentation In the period after the Italian artist engineers, from about 1700 to 1900, architecture increasingly separated from engineering. Two different points of view emerged—one emphasising aesthetic appearance (architecture) and the other the functional structure (engineering). One or the other tended to prevail—architectural for monumental buildings and engineering for utility buildings. In 1775 the Ecole des Ponts et Chaussées was founded in Paris, providing engineering education for the military. But there were also civilian needs and John Smeaton was the first to call himself a civil engineer. The gap between engineers and architects continued to grow ever wider in the nineteenth century through the steady demand for construction. Just as the masons had earlier formed guilds so people found benefit by developing specialised knowledge and skills and naturally, they wanted to get together to discuss mutual interests. Clubs, societies and professional institutions were set up—but the professions were fragmenting. The Institution of Civil Engineers was founded in 1818 and the Institute of British Architects in 1834. The burgeoning number of railway and manufacturing engineers wanted their own organisation, so the Institution of Mechanical Engineers followed in 1847. The pattern was repeated in America with the founding of the American Society of Civil Engineers in 1852 and the American Society of Mechanical Engineers in 1880. Ships were getting bigger and more complicated, so the Royal Institution of Naval Architects appeared in 1860. Naval architects are engineers—there is no distinct profession of architecture as there is for buildings. Naval architects are concerned with making ships structurally safe with a form that has a low resistance to flow through the water and which is stable at sea. The engineers who design and build mechanical and electrical equipment for ships are marine engineers and those who go to sea to maintain it are often referred to as naval engineers. By the end of the nineteenth century, the professions had become totally fragmented. Separate companies and partnerships of independent professionals emerged. Clients generally consulted the architects first to get design ideas. Consequently, where the engineers worked with architects, they were the junior partners. Architects such as Sir Gilbert Scott and Frank Lloyd Wright were amongst the first to ‘use’ engineers regularly. The architects needed engineers because the scientific and technical details of steel and reinforced concrete construction required technical training that most architects did not have. Any masonry-fronted structures with steel frames were sent to engineers—but well after the architectural concept design was formed. This often led to conflicts between the concept and the function—between the architecture and the engineering. In 1908 reinforced concrete was still a new material so the UK Concrete Institute was formed to exchange information and experience. The

52

4 Dwelling

first members were architects, engineers, chemists, manufacturers and surveyors. By 1912, the Institute began to embrace all areas of structural engineering, particularly steel frames. Structural engineering was defined as ‘that branch of engineering which deals with the scientific design, the construction and erection of structures of all kinds of material’. In 1922 the Institute became the Institution of Structural Engineers. In 2018 the Institution has worldwide influence through its members in 104 countries. By the end of the twentieth century construction engineering was so completely fragmented into sub-disciplines that major clients were becoming severely frustrated with inefficiencies, cost overruns and late delivery. They began to call for much more co-ordination and collaborative working. The principle of Preparedness had been lost through professional disjunctions and lack of coherent overview. The media usually assume that the architect is the ‘moving force’ behind a scheme and give little or no credit to the other professionals. In truth, clients have a variety of needs. Their decisions determine the balance of power in the building team. The way the various professionals interact is governed by the contractual arrangements and the nature of the end product. For example, if the client gives a very high priority to keeping a tight control on costs rather than architectural novelty, then he may be advised by his financial people to use a contractor-led design-build contract. In this type of arrangement, the architects and engineers usually play a subservient role. On the other hand, if a client wants a prestigious building with some architectural flourish then almost all structural engineers will provide a structural scheme for an architect-led design no matter how expensive. The question then becomes very basic: ‘How big is the client’s wallet?’ In truth, if there is no coming together between all of the requirements—especially the aesthetic and structural forms—then the solution is likely to be unsatisfactory. It is for these reasons that architects are now no longer automatically the leaders of the building team. But there can be a considerable downside—a lack of coherence with no clear allocations of responsibilities and accountabilities as may well prove to be the case for the tragedy of Grenfell Towers. The idea, mentioned earlier, of ‘The Engineer’ or ‘The Architect’ as a guiding mind3 or leader in overall charge of every detail of the works, has been drowned in the deep waters of the procuring, designing, building and maintaining of many commonplace buildings with very tight austere budgets. Whoever is appointed leader has to be appropriately qualified to protect the safety of the public and must be fully accountable for all decisions made or delegated—delegation within the building team does not imply abdication of duty of care. Of course, in practice, anyone taking on such heavy responsibility and culpability would need to be appropriately remunerated and protected in law from malicious claims—just as a surgeon who takes on responsibility for the life and death of his patient or a CEO takes responsibility for the performance of his company.

Architectural and Structural Form

53

Architectural and Structural Form The complexity of the technical detail of modern engineering practice lies behind the historical fragmentation of the professions and as such is entirely understandable. And it will continue, unless attitudes change, since the future holds even more possibilities for specialisation because IT promises robotic construction and techniques such as 3D printing—even with concrete.4 The architecture and engineering professions will continue to be concerned with the form and shape of a structure from different points of view. But there is a need now, more than ever, for the professions to come together, to collaborate right from the start. Collaboration depends on the principle of learning to bring differing perspectives together in harmony—to see the point of view of others. Architecture is important because it is about the sense of space, occupancy by people, symbolism, aesthetics and relationship to its setting. The artistic skills of the modern architects are special. But so are the skills of those who make a building work—the modern engineers. The consequences of getting the architectural form wrong are unsatisfied clients and public—no one will be injured or killed. The consequences of getting the structural form wrong are serious. If a building collapses people will be injured and possibly killed. If the engineering contributions continue to go unrecognised and taken for granted, then future recruitment of engineers will continue to be compromised. The media often attribute the whole of the credit to high profile ‘starchitects’. But without Arup engineering, Zaha Hadid’s sweeping contours of the Bridge Pavilion in Spain would not have left the drawing board. Architecture for buildings of high prestige of flagrant imagery such as Frank Gehry’s Guggenheim Museum in Bilbao (Fig. 4.5) has been criticised for showy, daring functionless forms that do not fit the locality. In that building, the engineering is completely subordinated to the architecture. The complex lines presented the eminent Indian American structural engineer Srinivasa Iyengar, a director of Skidmore Owings and Merrill, with significant challenges to make it stand up safely—yet how many people have heard of him outside of his profession? Andrew Saint of the Bartlett School of Architecture in London has written ‘In an art obsessed world, the architect had dragged the engineer out of the temple of reason and beguiled him to worship in the temple of art’.5

How to Read a Structure We can see and appreciate architecture—it’s there in front of you. Likewise, when you look at ‘naked’ structures like bridges and dams all you see is the structure—it’s as though you are looking at a human as a skeleton and not the whole body covered in flesh. Bridges have no ‘flesh’. A building is different—it has flesh and just as the human skeleton is hidden from view, so the structure is largely hidden within a building. You can see the structure as a building is being constructed before the exterior cladding is added—typically consisting of beams and columns.

54

4 Dwelling

Fig. 4.5 The Guggenheim Museum Bilbao Spain Architect Frank Gehry. Structural Engineers led by Srinivasa ‘Hal’ Iyengar of Skidmore Owings and Merrill. Image by PA. CC BY-SA-4.0 via Wikipedia Commons

The structural engineer’s job is to create the ‘skeleton and muscles’ that ensure that a structure stands firm no matter what Mother Nature comes up with. The demands are many and various—foreseen and unforeseen. Wind, earthquakes, the weight and activities of people, ground movements due to subsidence or settlements, road traffic, trains or crashes such as a railway accident or a ship docking out of control, the action of huge waves, intense storms, lightning or other natural events all cause forces on a structure. Some demands are legitimate such as people walking across a bridge with a weight limit, but others may be illegitimate such as those due to an overladen truck or car exceeding the speed limit. So, the engineer has to estimate the likely magnitude of these demands and how that demand relates to the strength of the structure. But how far should this planning go? Does it include tsunamis and terrorist attacks such as 9/11? There is no magical answer. The engineer applies common sense and makes a reasonable estimate. Nowadays there is help. Most countries write regulations as ‘codes of practice’ to give some uniformity of what it is reasonable to choose but the rules are often based on very sparse statistical data. The engineer has a duty of care to design and build the structure so that the resistance to those regulatory demands is greater by a suitable margin of safety. The chances of collapse must be acceptably low—but they are never zero because of the embedded uncertainty. One way of getting some idea about how a structure works internally is to take a piece of string and pull on it on both ends. Your pull at one end is balanced by your equal and opposite pull at the other. The string carries the flow of force from one end

How to Read a Structure

55

to the other. Your pull at one end represents the external demands that cause forces on a structure. Your pull on the other end represents the reaction of the ground. The internal force in the string represents the flow of internal forces. In the structure of a real building, the flow of forces is quite complicated. As long as nothing moves then the external forces and the internal forces balance in equilibrium. Just as your piece of string will snap if you pull too hard then any part of a building structure may break if the internal force effect gets to be too big. In general, there are three ways in which the materials of the structure are strong—pulling, pushing and sliding. Engineers use the term tension for pulling, compression for pushing and shear for sliding and they can occur in various combinations. The strength of the string is important but so is its flexibility—the amount of stretch—like a spring or elastic band. Engineers plot a graph of the force vs the amount of stretch or compression. If that graph turns out as a straight line, then the material is described as linear. If when the force is released the string returns to the original length, then we call it elastic (like an elastic band). In practice, few materials are linear elastic but often structures are designed as though they are to make the calculations tractable. This is one of the most common and important approximations of theory used in practice—though engineers always make sure that the approximations are made such that the structure is safe. The string structure is what engineers call ‘statically determinate’. All that means is that we can find the internal force purely through knowing the structure is in equilibrium. Many real structures are more complicated and can’t be solved this way because they are ‘statically indeterminate’. In other words, a state of equilibrium between internal and external forces is necessary and required, but it may not be sufficient or enough because there are too many constraints on movement. If we want to find the flow of all of the internal forces, we need to know more. The way engineers do this is to use three conditions and the energy ‘locked’ into the structure. The first condition, as we have already seen, is equilibrium. The second condition makes sure we represent the way in which a material responds to the internal forces—for example, how it stretches under tension—these are call constitutive relations. Third, various bits of the structure must remain properly connected, i.e. all displacements have to be compatible so bits of the structure (in the model) do not separate off. The energy locked-into the structure is potential energy, which as you may recall comes about because of the position of an object. So potential energy is created in (locked-in) our string when it stretches because the ends of the string move relative to each other. The same thing happens when some part of a full-scale structure is displaced. For example, a heavy weight will make every part of a structure, no matter how little or large, to move or deform very slightly. Consequently, all of the ends of the structural components move relative to each other and this creates potential energy in each one. That potential energy gets stored or locked-into the structure as the structure strains, so engineers call it strain energy. In more general engineering terms this strain energy is a form of inductance or accumulated flow of force—just as a water reservoir is an accumulated flow of water or a battery is an accumulated flow of electricity. It turns out that when a loaded structure is in equilibrium the strain energy is at a minimum.

56

4 Dwelling

These three conditions are satisfied along with energy calculations that are almost always now performed in a digital computer. The most well-known tool is ‘The Finite Element (FE) Method’. The methods, computer programmes and all the analysis behind them is ingenious theoretical engineering which started from esoteric research in the complicated mathematics of matrices (a matrix is an array of numbers or mathematical variables). The work started in the 1940s but was developed in the 1960s and 70s by John Argyris in Germany, Ray Clough in California and Olgierd Zienkiewicz in Swansea, Wales and many others. Now the tool is commonplace and available to almost all structural engineers. They use it to create a theoretical model of their structure on their own computers. The purpose of the model is to help the engineer understand the internal forces and displacements in the structure so she can make sensible and safe decisions about the materials, sizes and dimensions. The way this is done is to divide the total structure (in the theoretical model— not in reality) into discrete finite elements or pieces of structure—a bit like lots of interconnected theoretical Lego bricks. The bricks or finite elements may be any size but are usually either a whole structural member such as a beam or imaginary divisions (of a plate, wall or floor) into simple shapes such as triangles and rectangles. Like Lego bricks, the elements have to be connected to each other but not through studs into holes but by theoretical links at the nodes or corners of each element to the nodes of neighbouring elements. These link connections have to satisfy all of the three conditions locally. The engineer inputs data into the computer so it can calculate the internal and external forces acting at a node along a given direction. That direction has to be independent of all others, so it is called a degree of freedom—degree meaning amount, and freedom meaning the direction in which it is able to deform. Then these forces are equated. The equations include, as unknown variables to be calculated, the displacements along that degree of freedom. These are the constitutive equations that satisfy the second condition locally and also make sure there is local equilibrium, the first condition and local compatibility, the third condition. There is one mathematical equation for each of the 100 s of degrees of freedom over all of the links in the entire structure. The equations are assembled into matrices of simultaneous equations that together represent the whole structure. Simultaneous just means that the unknown values of displacements appear in more than one equation so the whole set have to be solved together. We humans can’t do that easily, but computers can. Once the computer has done its calculations to find all of the displacements along each degree of freedom then it can use those values to calculate all of the internal forces in the structure. This is the bare bones of how engineers like Sarah Buck and Michelle McDowell find the forces flowing internally within the structure of a building. Both have had remarkable careers. In 2007–9 Sarah was the 88th person but first woman elected President, of the Institution of Structural Engineers a great achievement for someone who was the first girl from her school to study engineering. After graduating she thought medical engineering research was for her. After a year she felt she needed to experience life in industry so took a series of jobs as a structural engineer. She branched out with her own business in 1993, and then three years later formed a company with a former colleague. Her work has included designing and supervising

How to Read a Structure

57

the construction of new Children’s Hospices, remedial works of an historic harbour, new and refurbishments of buildings such as schools, flats, offices, leisure centres, hotels, banks and airports—mostly in the UK but also in South Africa and Jamaica. She has an OBE for services to engineering and education. Like Sarah, Michelle McDowell went to an all-girls school (in Northern Ireland) where engineering was not on the radar. In common with many women of her generation, she found out about the profession by chance. She graduated in 1984 and now she has an MBE for services to construction. Then in 2011, she was the UK Veuve Clicquot Business Woman of the Year. One of the ideas that drives Michelle’s work is that she thinks that it is very important to understand other people’s viewpoints and to value and give credit for other people’s inputs—one of the reasons why she is a good leader. Her first job was with WS Atkins in Epsom, Surrey where she worked on a housing development in Oman and a major exhibition hall in Macau and then was seconded to John Laing Construction as a site engineer. She saw how good design can minimise risk and improve the building process. In 1989, Michelle moved to a company with a reputation for innovation with engineers and architects working closely together. She is amongst the few who give high practical importance to re-integrating the building professions—the principle of Preparedness in action. She worked on two notable projects in Berlin—the new Stock Exchange and the new British Embassy. The 9-storey superstructure of the Stock Exchange is suspended from concrete arches to create a large open space at ground floor. The British Embassy has an innovative winter garden and was designed with special security measures and proof against bomb blast. Michelle was promoted to associate partner but moved to BDP (Building Design Partnership) in 1997 as an Associate Director because she was attracted to their interdisciplinary approach. They employ engineers, architects, planners and designers in integrated teams. She was elected head of the company’s Civil and Structural Engineering group in 2004. One of Michelle’s important projects was the first comprehensive overhaul of the Royal Albert Hall since it opened in 1871. The project won the Europa Nostra Award in 2004—a prize that celebrates and promotes best practices related to heritage conservation, management, research, education and communication. The citation read: a ‘remarkable achievement of restoring and enhancing an international concert hall whilst still in use and for the ingenious provision of service requirements, thereby releasing space for the public’s enjoyment’. Michelle also led the building of the new HQ building for the pharmaceutical company Roche. The client saved so much money from the project that they were able to construct another building on the site without going over the budget or original timescale. The project won the British Council of Offices ‘Best of the Best’ award in 2006—the judges said, ‘Roche asked for quality without ostentation to attract and hold the best staff and this is what designer BDP has delivered’. In 2018 Michelle is leading the restoration and renewal of the Palace of Westminster for BDP with a fully multi-disciplinary team from architecture and engineering to planning, lighting, sustainability, acoustics etc. She says, ‘It really is a privilege to work on such an iconic structure and there is a lot to do’. Sarah and Michelle know that relatively simple demands such as the weight of people standing on a structure are straightforward to allow for. But Mother Nature is

58

4 Dwelling

ingenious. She will search out weaknesses and will test the firmness of our structures over years—if there is a weakness eventually, she will find it. Wind is perhaps the most obvious way she works. Wind is air in motion—caused by differences in air density and flow from high to low pressures. The Earth is rotating and so air does not flow in a straight line from high to low pressure but rather around the regions of high and low pressure as we see on TV weather forecast charts. The science of wind is complex and the effects on buildings and structures difficult to predict. The air which affects us on the surface of the earth is a layer seven or eight miles thick and called the troposphere. The steady winds and the various gusts that we experience depend on topography and factors such as the height above sea level and the pitch of sloping roofs, amongst other things.

Wind and Vibrations Until about the 1930s buildings were, by modern-day standards, relatively massive. The way in which engineers dealt with wind was consequently rather simple. The skyscraper building boom of the 1930s changed all that—wind became a much more important factor. Tall buildings started to sway backwards and forwards causing some discomfort to the people inside. The issues became more pressing for several reasons. For a start, structures had less mass because lighter materials were being used (largely a change from masonry to concrete and steel). They became less stiff (more flexible), which meant that they deflected more under a given load. And they also had less damping. Damping is the dissipation of energy, usually due to friction between constituent parts—like the shock absorbers of your car. A child’s swing reduces when you stop pushing because friction in the hinges dampens the motion until it stops altogether. Alfred Pugsley, you will recall, was a member of the Ronan Point tribunal. He was the one who identified the possibility that wind could have triggered the collapse by blowing out one of the panels. He was an engineering polymath who had extraordinary vision across a wide range of engineering matters because of his working experience in both aeronautical and civil engineering. His understanding of the effects of wind on building came from his early career experience working at Farnborough on the performance of aircraft wings. He worked on a phenomenon called flutter. Flutter6 is a kind of oscillation in the wind that can be self-reinforcing— engineers call it positive feedback. It happens when the wing or other structure, such as a bridge or building, develops large displacements as the flow of the wind causes vortices of air swirling around the structure and this can cause forces that were originally in alignment to be no longer in line (Fig. 4.6a). The out of alignment causes changes in the complicated interactions between the aerodynamic forces (exerted on the structure by the flow of air), the dynamic forces (because the structure is accelerating as it oscillates) and the response forces (because the structure deforms and the movement is damped). At a particular wind velocity called the flutter critical speed of the structure, the damping becomes insufficient

Wind and Vibrations

59

Fig. 4.6 Large displacements and flutter

and the displacements grow in cycles as in (Fig. 4.6b) and the structure becomes unstable. The result can be self-destruction. Flutter is not confined to aircraft wings unfortunately. The Tacoma Narrows Bridge that shook itself to pieces and collapsed in 1940 is often shown on TV as an example of what can happen. Alfred Pugsley’s work on flutter led directly to his student Alan Davenport taking up the baton. Alan was an exceptional engineer both technically and personally. He designed and built a wind tunnel laboratory in London, Ontario, Canada in which he analysed the effect of wind on a significant portion of the world’s tallest buildings and bridges including the Sears (Willis) Tower in Chicago, the World Trade Centre in New York in the 1970s and the Normandy Bridge, the Storebaelt bridge in Denmark and the Tsing Ma bridge in Hong Kong in the 1990s. Alan and his team analysed the wind flow and load over the structures in the wind tunnel and detected vulnerabilities which required changes in the design. Alan also contributed internationally to design standards. Alfred Pugsley also recruited a young applied mathematician, Roy Severn, who became one of the world’s leading earthquake engineers. Roy was appalled at the differences in damage from similarly severe earthquakes in very different countries. Earthquakes cause damage in well-prepared states and countries like California and New Zealand but the devastation in less well-developed countries like Pakistan and Iran is disproportionately greater—but could be moderated by good engineering. Roy’s research was three-pronged. First, he and his team developed new analytical calculations. Second, they installed a large ‘shaking table’ in their laboratory on which they could simulate ground motions. Third, they created a machine that could be bolted-on to actual full-scale structures to test them. The machine had rotating masses to excite and measure how the vibrations are transmitted through the structure. In 1990, Roy was elected President of the Institution of Civil Engineers—one of the highest honours in the profession.

60

4 Dwelling

Earthquakes can destroy structures in large areas and bury people under piles of concrete rubble. They are the result of rock deep inside the earth slipping along a fault line and causing shock waves to travel through the ground and along the surface just like waves in water. The motions can cause roof beams to slip off supports and rectangular unbraced storeys to lozenge and collapse. Good engineering can help structures sway and absorb the flow of force and energy just as you might sway and absorb the motions as you walk on the deck of a ship. Very tall structures often contain energy absorbers such as a large mass or frictional spring. The dynamic response of sensitive structures such as skyscrapers, large dams holding back huge volumes of water and big bridges are complex and as recent earthquakes have shown there is still much to learn.

Progressive Collapse and Terrorism The failure of Ronan Point was traumatic for the profession—but more was yet to come and the learning about the vulnerabilities of skyscrapers to progressive collapse continues. After Ronan Point, the reaction outside of the UK to the possibility of progressive collapse was markedly less. For example, only a brief statement was included in an American National Standard for buildings in 1972. The authorities called for it to be considered—but there was no development or discussion. At that time the problem was not well understood though some researchers began to investigate. Then, in 1995, a second wave of interest followed for very different reasons—terrorism. An attack on the Alfred P. Murrah Federal Building in Oklahoma killed 168 people (Fig. 4.7). The trigger for the event—a terrorist truck bomb—was of course quite different to that at Ronan Point but one that has unfortunately become all too

Fig. 4.7 a Alfred P. Murrah Building Oklahoma City, USA before attack in 1995. b Alfred P. Murrah Building Oklahoma City, USA after attack in 1995. By permission of the Oklahoma City National Memorial & Museum, USA

Progressive Collapse and Terrorism

61

common. Then came 9/11 and the destruction of the twin towers of the World Trade Centre in New York in 2001. The structures of both of these buildings were quite unlike Ronan Point or each other. The Murrah building was a 9-storey office with a conventional reinforced concrete frame of beams and columns in a largely rectangular grid. However, it had one critical and relatively unusual feature. Running along the front of the building was a large horizontal girder supporting the rest of the building, and itself supported on a smaller number of ground floor to first floor columns, to allow easier access to the building. The truck bomb parked outside the entrance took out some of the columns supporting the big girder and consequently about half of the floor area of the building collapsed. The New York World Trade Centre Twin Towers were designed in the late 1960s and early 1970s as a new kind of lighter weight structure using modular construction (Fig. 4.8). Leslie Robertson was the lead structural engineer. After graduating in civil engineering Leslie worked in Seattle for Worthington, Skilling, Helle and Jackson (WSHJ) and became a partner in 1967. In 1982 he formed his own practice. Over his career, he has won many awards and prizes for his work on high-rise buildings. The idea of a skyscraper as a gigantic vertical tube had been developed earlier by Fazlur Khan an American structural engineer born in Bangladesh. Fazlur first qualified in Bangladesh then took two master’s degrees and a Ph.D. in the USA before becoming a

Fig. 4.8 The World Trade Centre, New York before it was brought down on 9/11. Image by MesserWoland, Public domain via Wikipedia Commons

62

4 Dwelling

US citizen. He took a job with Skidmore, Owings & Merrill in Chicago and was made a partner in 1966. He was at the forefront of a renaissance in skyscraper buildings in the USA when he led the design of the 100-storey John Hancock Building in Chicago 1968 and the 110-storey Sears Tower (later renamed the Willis Tower) opened in 1973 also in Chicago. Leslie Robertson chose tubes with sides of 64 m made up of closely connected smaller steel tubes or boxes for the ill-fated WTC in New York. Inside the main tube was a similarly constructed central core that contained the lifts, staircases and utilities. The core and the perimeter were connected at each of the 110 floors by beam joists supporting concrete slabs. The impact of the terrorist-directed planes caused huge fires. As a result, floors collapsed onto lowers floors, the outer perimeter boxes began to bow, and the floors began to cascade down the building in domino-style. The net effect was that the building did not topple but fell more or less on its own plan area. Just a year after those tragic events I met Leslie Robertson in his office overlooking the site where the twin towers had been. I asked him how he was feeling after the intense publicity. His reply was simple ‘I would give anything not to be the most famous structural engineer in the world right now’. Leslie showed me plans for the Shanghai World Financial Centre building that was being engineered in his office. They were very impressive. At that time the plan was to build over 100 storeys with a gigantic circular hole right through the building near the top. Not only that they planned a ride similar to the London ‘Eye’ within the hole. The ride and the views would have been spectacular. Many years later in 2010 my wife and I visited Shanghai and saw the completed building. The idea of the circular hole had been abandoned and replaced with a rectangular one and consequently no ride (Fig. 4.9). Nevertheless, the building itself is breath-taking. Our Chinese hosts took us up to the observational tower and we walked on the glass floor with some trepidation some 474 m. in the air. It was a comfort to know that guarding against progressive collapse is now at the forefront of the thinking of structural engineering designers.

Structural Safety Safety is paramount. Once you erect a pile of steel and concrete up in the air then if you don’t get it right Mother Nature will find you out and bring it down and probably kill people. Safety is consequently the subject of much research and debate. Structural engineers of such enormous skyscrapers have to have confidence in their ability to deliver safe buildings. So, should the phenomena of progressive collapse of buildings have been anticipated? At one level the answer is straightforward—yes. But hindsight is a wonderful thing. I used the analogy with the house of cards for Ronan Point to illustrate a point about faults in the basic structural design of that building—but the real event was rather more subtle, as the report of the full inquiry showed. The faults in the design of the cladding at Grenfell remain to be discovered. The reasons for

Structural Safety

63

Fig. 4.9 The Shanghai International Finance Centre. Image by GG001213. CC0 1.0 Universal Public domain dedication via Wikipedia Commons

64

4 Dwelling

Fig. 4.10 Professor Sir Alfred Pugsley (1903–1998). By permission of the Special Collections of the University of Bristol

the collapse of the Murrah building are a little less clear, and the way the WTC failed was not really anticipated by anyone. In any case, both American events were triggered by totally unforeseen acts of terrorism. These events help us to get a feel for the central dilemma for engineers—how can we predict what might happen to the things we make? In an uncertain world, how can we ensure that they remain safe? In this new age of complexity, we can’t simply assume the future will be like the past—surprises will come to haunt us. We have a duty to take the principle of the Unintended very seriously and combine it with the principle of Preparedness. Alfred Pugsley was probably appointed to the Ronan Point tribunal because he was a world-leading expert in the engineering of structural safety. His interdisciplinary experience in aircraft, buildings and bridges gave him and an unusually wide perspective that allowed him to cross fertilise ideas. Each of these structures are engineered differently but the underlying theory is the same. His influence was also in large part extended by his ability to attract some of the most able postgraduate students, such as Alan Davenport, Lord Henry Chilver, John Caldwell and Tony Flint, who all went on to become some of the UK’s most successful engineers (Fig. 4.10). Alfred

Structural Safety

65

Pugsley’s interest in safety was sparked early in his career also through direct practical experience. In the late 1920s, he worked at the Royal Airship Works, Cardington, UK as a technical officer on the building of Airship R101—his engineering experience of airships informed his later contributions to civil engineering. Unfortunately, the R101 crashed on its maiden flight to India in 1930 for a complex set of reasons—but the final trigger was probably a tear causing a gas bag to fail. The inquiry into that accident concluded that public policy had exerted pressure to undertake the journey—in other words, the ship flew before it was really ready. Lord Thomson, the Air Minister and many of the designers were killed and the crash effectively ended British involvement in airships. As a young engineer, Alfred Pugsley must have been deeply moved. He took a job at the Royal Aircraft Establishment, Farnborough, UK where one of his fellow workers there was the novelist Nevil Shute. Shute, we are told, developed one of his most famous stories after Alfred Pugsley sent him some technical papers about metal fatigue. Theodore Honey in the novel No Highway8 is the engineering scientist doing the work similar to that which Alfred Pugsley had done. However, the personal characteristics of Alfred Pugsley and his fictional counterpart were strikingly different. Fatigue of metal is a complex phenomenon and cannot be predicted with total accuracy. However, engineers have sufficient understanding to keep it under control—as long as they are diligent. Alfred Pugsley would certainly not have agreed with Theodore Honey’s confidence in his calculations—and neither would Alfred Freudenthal an eminent American engineer also renowned for his seminal work on metal fatigue and structural safety. The need for close control is one of the reasons why aircraft have to be constantly monitored and maintained—looking for possible cracks. Fatigue occurs in all structures, but particularly in aeroplanes and bridges, because they are subject to many cycles of repeated loading. Nevil Shute should have known better because he worked as an aeronautical engineer for de Havilland, the company that produced the world’s first commercial jet airliner—the 106 Comet with a maiden flight in 1949. A year after entering commercial service three Comets broke up in mid-flight. There was metal fatigue in the airframe structure. Over its whole life, there were 13 fatal Comet crashes and 426 people killed. The structure was modified but sales never really recovered. However, the Comet 4 of 1958 flew for 30 years and the Hawker Siddeley Nimrod, a maritime patrol aircraft developed from the Comet, remained in service with the RAF until 2011. One of the important legacies of Alfred Pugsley’s work is that in a fragmented profession it is important to learn lessons across disciplines. Through his direct experience and influence lessons from the aeronautical engineering of R101 were passed onto civil engineers. He was acutely aware of the dangers of fragmentation and the need for interdisciplinary working. Alfred Freudenthal was Alfred Pugsley’s contemporary with a parallel but different career. He was born in what is now Poland in 1906. He studied civil engineering in Prague but moved to Palestine to work as a senior engineer constructing a port in Tel Aviv. He took an academic post in Haifa before moving to the USA and eventually a professor at Columbia and the George Washington Universities. He pioneered the use

66

4 Dwelling

Fig. 4.11 ASCE Alfred Freudenthal Medal. With permission from the American Society of Civil Engineers (ASCE). For internet posting, add the following additional notice: “This material may be downloaded for personal use only. Any other use requires prior permission of the American Society of Civil Engineers. This material may be found at [URL/link of abstract in the ASCE Library or Civil Engineering Database]”

of statistics and probability theory in the safety of structures. In 1975 the American Society for Civil Engineers donated the Freudenthal medal in his honour (Fig. 4.11). Alfred Freudenthal’s work was taken up by engineers who went on to become major figures in researching structural safety such as Masanobu Shinozuka, Jack Benjamin and Allin Cornell.9 Although their research work is of first-class academic quality, it was not informed by Alfred Pugsley’s wider thinking or by the principles of the Unintended, or of Preparedness. Structural safety was formulated as a problem of probability and statistics where the principal task is to predict the chances of failure given a deterministic model of a system with parameters that are modelled as random variables. This approach is fine if there is little systemic change but that is not so for complex systems—we cannot assume that data from the past will necessarily reflect conditions in the future. For example, the return periods for extreme weather events will get shorter. The statistical approach does not give sufficient attention to uncertainties in underlying deterministic models and processes to manage the consequences of inevitable human frailties. Good design, diligent control, attention both to detail and the ‘big picture’ overview, and active learning are essential parts of the construction process not normally included in probabilistic safety analyses. Consequently, the methods are useful, and influential but also partial because they omit many of the actual reasons why structures fail—such as Ronan Point and Grenfell Towers. Earthquakes, wind load, progressive collapse and earthquakes are just four of the complex natural phenomena that demonstrate how engineers have constantly to be ‘on guard’—striving to understand and control the behaviour of the things they make and actively using the principle of Preparedness. In these four cases, even to this day,

Structural Safety

67

the science is seriously incomplete. Part of being prepared is holding the principle of the Unintended at the forefront of our minds. That is why Nevil Shute’s implication that such factors can be predicted precisely is seriously misleading. But because there are things, we have yet to understand sufficiently well to predict them does not mean that the products of engineering cannot be managed safely—that is part of the art of engineering. There are many other phenomena of this type—for example, the behaviour of soils and rocks that support the foundations of structures, the creep and shrinkage of common materials like concrete and the steel inside nuclear reactors, the effects of a Mexican wave on the stability of a sports stadium, the effects of earthquakes on large dams, and the non-linear behaviour of very slim components such as turbine blades. The list of known unknowns is actually quite long—quite apart from the possibilities of unknown unknowns. To reinforce this point, we should note that over the centuries we have engineered many things before we had a good understanding of how they work. One of the most well-known, as we discover in the next chapter, was the building of steam engines long before there was any science of thermodynamics. There are risks in whatever we do. We have no choice but to acknowledge those risks and work diligently to keep them under control. That isn’t easy because engineers operate in the real world of messy compromise—not under precisely controlled laboratory conditions of the scientific experiment. So, some factors we may know a little about, but others may be things about which we know nothing. That is why we need people who are competent with practical wisdom and deep understanding to look after our big structures and cope with new surprises that our changing climate may throw at us. It follows that one of the biggest risks to our future dwellings and infrastructure is not only technical and narrow—it is also human, social and wide. For example, the differential impacts of earthquakes in various parts of the world are not narrowly technical but due to lack of understanding by political leaders who have insufficient ingenuity to see that the costs of engineering their infrastructure properly in the first place are less than dealing with the massive consequences of an event. In the Western world, we face a serious shortfall in competent, highly trained people. One of the reasons is that our collective understanding, and media reporting, of the nature of engineering, is inadequate and often misleading. We need young people, and particularly young women, to see engineering as the interesting and rewarding career it actually is—engineering by people for people. Buildings do not exist in isolation but are interconnected by road, rail and air—the engineering of moving. End Notes 1. Building a Safer Future Independent Review of Building Regulations and Fire Safety: Interim Report Presented to Parliament by the Secretary of State for Communities and Local Government by Command of Her Majesty December 2017 See https://www.gov.uk/government/uploads/system/uploads/attachment_ data/file/668831/Independent_Review_of_Building_Regulations_and_Fire_ Safety_web_accessible.pdf (Last accessed February 2019).

68

4 Dwelling

2. Turner, BA., Pidgeon, NF. (2nd edition 1997), Man Made Disasters ButterworthHeinemann Ltd, London. 3. A Guiding mind is a team leader with vision who inspires, promotes and checks out a project or situation—someone in charge, trustworthy and reliable who can hold the ‘big picture’ whilst mastering the detail so that all is integrated and coherent. 4. Richardson, V. (2017), 3D printing becomes concrete: exploring the structural potential of concrete 3D printing, The Structural Engineer. 95, 10 See also https://www.dezeen.com/2017/10/27/worlds-first-3d-printed-concretebridge-bicycles-bam-infra-netherlands/ (Last accessed February 2019). 5. Saint, A. (2008) Architect and Engineer: A Study in Sibling Rivalry Yale University Press. 6. Flutter is a type of wind excited unstable oscillation due to interactions between forces on a structure. Air flow causes a wing or bridge deck to move but stiffness and damping or friction oppose that movement—but if they are insufficient then the motions will increase and cause failure. 7. Quoted by Li Hu of Steven Holl Architects at http://www.nytimes.com/2006/05/ 21/magazine/21bejing.html?_r=1&oref=slogin (Last accessed February 2019). 8. No Highway is a good yarn but as much a human story about the ‘boffin’ Mr Honey as about the engineering and science etc. Nevil Shute’s biographer John Anderson (Parallel Motion published by Paper Tiger in 2011) quotes Shute on the subject: “You think that was my own idea? Look I’m getting a little embarrassed about being hailed a prophet of metal fatigue. It really happened this way. Someone sent me a couple of technical papers by Professor Pugsley and he forecast the whole thing. I thought it was a fascinating idea for a novel so I wrote it. If anyone was the prophet for that book it was the Prof.” 9. The first of a series of International Conferences on Structural Safety And Reliability (ICOSSAR) were initiated by Alfred Freudenthal in 1957.

Chapter 5

Moving

Suck, squeeze, bang, blow, turbine wheels that make it go

Edith’s Story In 1898, Edith Stephenson (not her real name) was cycling through the English countryside on her own. She was enjoying the new-found freedom of her ‘safety’ bicycle. Women riders liked these new machines because they were much easier to ride in female attire and much less risky than the alternative—the penny-farthing. The large wheel of the penny-farthing was popular with young men showing off how fast they could go. The youthful risk takers on their ‘boneshakers’ (the name given to all cycles with an extremely uncomfortable rise caused by a stiff frame and hard wheels) could reach high speeds as they rattled over ruts, cobbles and stones rather like the skateboarders of today. The high riding position was daunting but could be mastered with practice and perseverance. The problem was that once mounted—nearly over the front axle—riders could pitch forward and tumble off head first. Accidents were quite common and sometimes even fatal. Some cyclists relished coasting down hills with their feet off the pedals and over the handlebars—so they were pitched off feet first. Edith’s safety bicycle, unlike the penny-farthing, resembled a modern one. At the time most riders thought it a big improvement. For a start the two wheels were of a similar size so Edith could reach the ground with her feet and could stop more easily. Her pedals acting as a lever turned the rear wheel through a chain pulley. A large front sprocket, or toothed chain wheel, turned by the pedals was connected to a smaller sprocket on the rear wheel. In that way, each turn of the pedals was multiplied up. That was why the wheels could be smaller for similar speeds. Pneumatic tyres had replaced the previously solid ones—making for a smoother ride. The cycle frame © Springer Nature Switzerland AG 2020 D. Blockley, Creativity, Problem Solving, and Aesthetics in Engineering, https://doi.org/10.1007/978-3-030-38257-5_5

69

70

5 Moving

was in the form of two triangles, but some had a step-through frame made especially for women. The centre of gravity of the whole bicycle was much lower than the penny-farthing, so the danger of pitching over the handlebars was much less. In the 1890s, the safety bicycle created a bike boom. Edith was dressed in knickerbockers—that was probably the most practical and comfortable clothing at the time for a female cyclist. The story goes that as she rode, she eventually came to an inn and decided she would like to take some refreshment. She was somewhat shocked when the innkeeper refused her a seat in the coffee room. He insisted that to get service she would have to go into the public bar. Apparently, he objected to Edith’s clothes. He didn’t think it proper for a woman to appear in public in anything but a long skirt. Edith objected but to no avail. Indeed, she was incensed and later took her case to court—but she lost. The court ruled that the innkeeper had a right to refuse service. The incident was later used1 as an example in the battle for equal rights for women. The bicycle had begun to play a part in challenging gender roles and to emancipate women. Susan B. Anthony, the American civil rights leader, said in 1896, ‘The bicycle has done more for the emancipation of women than anything else in the world’2 .

Moving Edith was cycling to enjoy the freedom of exploring the countryside—to see the world afresh. Travel gets us away from routines and allows us to re-create, meet new people, see new places and explore different lands. Explorers and settlers push back the frontiers and find new places to live and work. Moving ‘things’ is another of our basic human needs—we move ourselves, goods such as food, water and all kinds of products but also utilities such as gas, electricity and oil. The first question for this chapter is this. How have we engineered the ‘things’ that enable us to travel at ever-increasing speeds? For example, how did the bicycle evolve? The story demonstrates how the ideas of a few individuals can be influential in triggering yet more ideas—incremental development into an explosion of innovation. How have we learned to travel on the water—rivers, lakes, seas and oceans—to satisfy our sense of adventure, to explore and settle in other parts of the world by ‘fair means or foul’. One of the curiosities we find is that it took us a long time to understand that we can use heat to power machines—then quite quickly along came the steam engine and out of that, the internal combustion engine, the turbojet and the space rocket. A key to that further development was an understanding of the essential technical differences between solids and fluids. Our relatively recent discoveries of how to generate and use electricity are becoming even more important as we enter the age of the electric vehicle. Moving requires energy, so the engineering of moving is about managing the way energy flows. We coach natural forces to do what we want. All processes have energy as a potential drives a flow. For example, volts are the potential that drives a flow of amps and velocity is the potential that drives the flow of force. Potential

Moving

71

times flow is the power of a process and the rate at which energy is transferred and work is done. What is more whilst the parts of a thing may be immobile left on their own, when they collaborate adeptly, they can move—the principle of Part. For example, none of the subsystems of your body (muscles, skeleton, blood, digestion, etc.) can walk and talk on their own. But working together as one person—you—they can. The characteristics of this kind of collaborative working are known as emergent properties. They are characteristics of a whole but not of the parts—they emerge as properties of the whole because the parts work well together. In the same way, the parts of an engine work together to create the emergent properties of turning the wheels of your car or moving an aeroplane through the air. When emergent properties are well created, they are an aspect of the elegance of function which are manifest in the way potentials create the rhythms of energy flows and when discernibly appreciated are part of the creative artfulness—the aesthetic of function that I referred to in Chap. 2. We human beings have been so successfully ingenious that the walking and running of cavemen have relentlessly evolved, in ‘fits and starts’ with many ‘dead ends’, to today’s complex transportation systems. Many of us can travel to just about anywhere on the globe as we wish and can afford—but of course there are many who are not as fortunate. Engineering progress has come at a price—just think of traffic congestion, pollution and the stresses of negotiating airports. But transportation is set to change even more—driverless cars controlled by artificial intelligence are just one example which has become almost daily news. Being prepared and coping with these issues and their unintended consequences as they unfold will require all of our five principles. But to use them wisely we should appreciate what has gone before.

Early Land Transport The ancients had only their own energy—walking and running. Cave dwellers and hunter-gatherers roamed to search for food and shelter—following the seasons. They dragged large objects on a platform with smoothed runners—a sledge. It wasn’t long before they used domesticated animals such as dogs and horses to pull them. Solid wood wheels date from around 3,500 BCE—about the same time as the first horse riding. Similar wheels are still used even today in some remote parts of the world. Castrated cattle or oxen were easier to control and hence used to pull wagons and carts. Archaeological evidence suggests that wheel spokes were developed by around 2,500 BCE—presumably to lighten the heavy solid wheels and use components that were easier to handle. These early radial members were made by carving a log using a spoke-shave tool. An iron age spoked wheel dated around 1,000 BCE has been excavated at Choqa Zanbil in present-day Iran. Horse-drawn war chariots with spoked wheels were used for three centuries down to classical Greece. Celtic chariots had an iron rim around the wheel in the 1st millennium BCE. For a long time, there was little change so that even by the time of Tudor England you would do well to travel 50 km in a day by land. By the mid-seventeenth century, stagecoaches were

72

5 Moving

running regularly between major towns in England. The best way to travel long distances was on the water.

Water Transport The earliest boats were dugouts or hollowed-out tree trunks in about 8,000 BCE. Logs and papyrus reeds were tied together to make rafts and reed boats propelled by handheld paddles and poles. To get greater leverage, paddles developed into oars that are connected to the boats through holes or rowlocks or tholes (vertical wooden pegs). The oar is one of the earliest examples of an engine—simply a lever that magnifies the distance moved by the blade compared to that of the oarsman. The oar propels the boat further for each stroke in proportion to the lengths from the rowlocks. A lever is an elementary machine that increases the force of our hands—as any criminal using a jemmy knows. Planks attached to the sides of a dugout and braced across the width were the first signs of a boat hull. The precise dates of the first sails are uncertain but around 3,400 BCE boats were sailing on the River Nile. The wind blew mainly North to South, so they sailed upstream and used oars to travel downstream. By about 1,100 BCE the Phoenicians developed squat vessels rounded at both ends for carrying goods and passengers and a longer boat with a sharp bow which they used as a battering ram in war. Oars were preferred to sail to get extra speed and manoeuvrability. The best way of increasing speed was to add more oarsmen—this they did by adding an extra bank to make a bireme and then a third to make a trireme. Even quinquerimes, with five banks of oarsmen, were made by around 250 BCE. The first sails were square rig then came fore and aft rigs more suitable for rivers and estuaries. Later complicated rigs were developed for single mast cutters, double mast ketches and yawls and schooners with two or more masts. The golden age of sail from sixteenth to mid-nineteenth century was a period when international trade and naval warfare were dominated by sailing ships. Canals played a major role in the industrial revolution in England in the eighteenth century. One of the first, the Bridgewater Canal, was built by James Brindley to carry coal the 41 miles from Worsley to Manchester in England. Within a few years, a national network of canals emerged with the period 1770–1830 referred to by some as the golden age of canals in England. Canals are man-made waterways and built to be as level as possible. The engineering challenge is to cope with the inevitable differences in elevation. The answer is to build a series of stepped level sections linked by locks. A lock is a chamber with gates at both ends. When both gates are closed the level of the water in the lock can be changed to match the water level at either end by letting water in or out. Canal routes had to be planned and surveyed to be as short and level as possible. Shifting large amounts of earth was difficult and expensive so the survey had to include the planning of the ‘cut and fill’. The ‘cut’ is the amount of material that is excavated because ground level is too high and then transported and used to ‘fill’ areas where ground level is too low. If the total cut

Water Transport

73

matches the total fill, then there is no waste material to be dumped or extra material to be transported in. Another major challenge is to make sure the canals do not leak. The main early waterproof lining was ‘puddle’—a mixture of sand and clay. Modern materials may be concrete, fly ash (a coal combustion product), bentonite (a type of clay), bituminous materials and plastic sheeting. Sometimes cut and fill wasn’t enough, so canals had to be taken by aqueducts over deep valleys and through hills by tunnels. Perhaps the most famous and impressive historic aqueduct is at Pontcysyllte where the Llangollen Canal crosses the River Dee in North East Wales (Fig. 5.1). The spectacular 18 stone arches with a cast iron casing carrying the waterway are a well-known tourist attraction. The bridge was designed and built by Thomas Telford and William Jessop and opened in 1805. The Trent and Mersey Canal at Harecastle in Staffordshire goes through two tunnels to carry coal to the Staffordshire potteries. The first was built by James Brindley around 1777—he died before it was opened— and the second one by Thomas Telford in 1827. Other engineering challenges are the building cofferdams to create dry construction below the waterline to build bridge piers and foundations. Eventually, the railways took over and many of the UK canals fell into disrepair. But in recent years they have been renovated, often by volunteers. Now thousands of boats and millions of people enjoy vacations cruising and exploring the canals and enjoying the peace and tranquillity of the waterways.

Fig. 5.1 Pontcysyllte aqueduct

74

5 Moving

Land Transport Strangely it was some 5,000 years after the first use of wheels before Edith could mount her bicycle. The cycle was one of the first of the modern modes of land transport that developed from the nineteenth century. The story of the bicycle is of a purposeful pursuit of better human-powered transport. The timeline is incremental and cumulative because each stage of the story built on what had gone before. It is irregular because the time from the first wheel to someone having the idea to put two wheels in-line was considerable. But once that idea took hold only about 200 years or so elapsed between the earliest cycles and the ones we have today. And there was a dead end on the way—the penny-farthing. The ingenuity required at each stage to evolve the different mechanisms is not obvious in hindsight—a bicycle seems straightforward to us today—but actually lots of different people contributed to create the lever crank mechanism, gearing, pneumatic tyres and ball bearings, and in more recent times, new materials, disc brakes, tubeless tyres and electric motors. The timeline is powerful because we can understand each development without any specialist technical knowledge. All of us can see how we collectively engineer inventions, unlike later developments of the steam and internal combustion engines which are much more opaque to the non-specialist. The bicycle story began around 1817 when German Karl von Drais had the idea of building his draisine (dandy horse). Figure 5.23 shows the compelling timeline for the inexorable progress from the primitive draisine to the sophisticated lightweight machine on which Chris Froome won his fourth Tour de France in 2017. Two of the first improvements on the hobby horse were by Denis Johnson of London in 1818 and then in 1839 by Scottish blacksmith Kirkpatrick Macmillan who made the first treadle or foot lever. The rider pushed on two treadles alternatively. They were connected by long rods to a cranked bar on the rear axle and so was the first rear-wheel-driven bicycle. In 1840, Englishman William Sawyer of Dover invented a four-wheel velocipede with foot treadles as did Scotsman Gavin Dalzell in 1845. A German Karl Kech or a Frenchmen possibly Pierre Michaux and his son Ernest Michaux or Pierre Lallemant fixed pedals to the front wheels of a bicycle around 1864—they also called it a velocipede. Micaux et Cie started mass-producing bicycles in 1868 with a frame of two pieces of cast iron bolted together rather than wood. Scot Thomas McCall built a treadle bicycle in 1869 and in the same year the New York Company Pickering and Davis developed a pedal bicycle for ladies. Frenchman Eugene Meyer invented the wire-spoke tension wheel. A spoke is a rod radiating from the hub of a wheel to the outer rim. The wooden or iron-spoked wheel was used without much modification until the 1870s when wire wheels and rubber tyres were invented. In a simple wooden wheel, the load on the hub causes the wheel rim to flatten slightly against the ground as the bottom spoke shortens and compresses. The other wooden spokes don’t change. They were used for horse-drawn carriages and wagons and early motor cars. By the nineteenth century, the heavy wooden-spoked wheels were replaced by lighter wheels for bicycles, wheelchairs, motorcycles and cars. These new wheels had tensioned, adjustable metal wires as

Land Transport

75

Figure

Date

Title

5.2.1

3,500

The earliest wheel The wheel

Innovation

BCE

was solid wood. The Sumerians in Mesopotamia built a twowheeled chariot for transport and war, sometime between 3500 and 3000 BCE.

5.2.2

1420

Bellicorum

A man

instrumentorum

powered

liber, cum figuris

vehicle

et fictitys litoris conscriptus (Illustrated book of war instruments). by Venetian Giovanni de la Fontana (1395 – 1455) includes a drawing of a four wheeled velocipede powered by rope connected by gears. 5.2.3

1493

A replica of a

If not a forgery

velocipede was

this was the

made in 1965 – 72 first example based on a sketch

of a man-

said to have been

powered

drawn by a pupil

vehicle with

of Leonardo da

two wheels in

Vinci. However,

line.

the sketch may have been a forgery

Fig. 5.2 A timeline of the bicycle

Image

76

5 Moving Figure

Date

Title

Innovation

5.2.4

1817

Karl von Drais

The first

developed the

known human

German draisine

powered

(running machine, vehicle with dandy-horse or

two in-line

velocipede).

wheels. The rider pushed against the ground and could use a chest pad to create extra thrust.

5.2.5

5.2.6

1818

1818

Denis Johnson of

Use of iron

London improved

rather than

the dandy horse.

wood.

Denis Johnson’s

Could be

steerable dandy

steered and

horse for women.

was easier for women to mount.

5.2.7

1839

Scottish

Long push rods

blacksmith

were connected

Kirkpatrick

to crank rods

Macmillan made

on the rear

the first treadle

axle.

powered machine.

Fig. 5.2 (continued)

Image

Land Transport

77

Figure

Date

Title

Innovation

5.2.8

1840

Englishman

Four wheels

William Sawyer

were more

of Dover made a

stable than two.

four-wheel velocipede with foot treadles. He presented one to the Prince of Wales in 1858

5.2.9

1864

A German Karl

One of the first

Kech or

examples of

Frenchmen Pierre

using pedals to

Michaux and his

turn the front

son Ernest or

wheel.

Pierre Lallemant attached pedals to the front wheels. 5.2.10

1868

A company called

Perhaps the

Micaux et Cie

first example

started mass

of large scale

producing

(mass)

bicycles with a

production.

frame of two pieces of cast iron bolted together.

5.2.11

1869

New York

Designed

Company

especially for

Pickering and

ladies with

Davis developed

long skirts to

a pedal bicycle

mount

for women.

Fig. 5.2 (continued)

Image

78

5 Moving Figure

Date

Title

Innovation

5.2.12

1869

Frenchman

Wire spokes in

Eugene Meyer

tension. Large

invented the wire

front wheel to

spoke wheel and

go faster

made a ‘penny farthing’ with a large front wheel.

5.2.13

1886

A Coventry

In 1820s–1850s

Rotary

a variety of 3

Quadracycle

and 4 wheelers

tandem for two.

of different

The White House, designs with Washington DC

pedals,

is in the

treadles, and

background.

hand cranks were tried but they were heavy with high rolling resistance

5.2.14

1879

Englishman

An early

Henry Lawson

example of

built a

pedals and

‘bicyclette’.

chain that drove a rear wheel.

1880

Cup and cone

Reduced

ball-bearings used frictional in bicycles in

resistance in

USA

the moving parts.

Fig. 5.2 (continued)

Image

Land Transport

79

Figure

Date

Title

Innovation

5.2.15

1884

The English

Pedals and

Coventry based

gearing

company of

through a chain

Hillman, Herbert

drive onto the

and Cooper

front wheel.

produced the

Thought to be

Kangaroo bicycle. safer and easier It was probably

to ride than the

the first to be

penny farthing.

called a ‘safety bicycle’. 5.2.16

5.2.17

1884

1884

John McCammon

Step through

of Belfast

frame to allow

developed a

easier

safety bicycle.

mounting

Englishman

Frame was

Thomas Humber

recognisably

built a safety

similar to a

bicycle. The

modern

image is from

bicycycle.

1890

5.2.18

About

Mid nineteenth

Tricycles

1885

century ‘Tricycle

became

de Léopold II’ of

popular.

Belgium. It is said he rode it to see his secret teenage French lover.

Fig. 5.2 (continued)

Image

80

5 Moving Figure

Date

Title

Innovation

5.2.19

1885

The Whippet

Two speed

safety bicycle.

derailleur type gears to the rear wheel. Derailleur is French for derailment of train – presumably indicating movement of chain between gear wheels

5.2.20

1885

Englishman John

The pneumatic

Starley developed

tyre was

a safety bicycle

developed for

His employee

this safety

George Franks

bicycle.

suggested the name Rover – later famous for its motorcycles and then cars 5.2.21

1887

Scottish born

The pneumatic

John Dunlop

tyre made a

developed the

more

pneumatic tyre.

comfortable ride but was difficult to repair.

Fig. 5.2 (continued)

Image

Land Transport

81

Figure

Date

Title

Innovation

5.2.22

1891

Charles Terront

Frenchman

on the front cover

Edouard

of the Petit

Michelin

Journal on his

produced a

way to winning

detachable tyre

the Paris-Brest-

for bicycles

Paris cycle race

and so

with removable

puncture

Michelin tyres.

repairs were much easier.

5.2.23

1893

The London

The fore runner

Crypto Cycle

of modern 3

Company

speed gear -

produced the

now usually on

Alpha Bantam

the rear wheel.

with a selfcontained gear in the hub of the front wheel.

Fig. 5.2 (continued)

Image

82

5 Moving Figure

Date

Title

Innovation

5.2.24

1897

Earliest folding

First folding

bicycles

bicycles. A

developed for the

Pederson

bicycle infantry.

folding bicycle

The image is at

was used in the

Fort Missoula,

Second Boer

USA.

War (18991902).

5.2.25

1917

Italian Bersaglieri

Improved

(light infantry

mobility for

troops) before

infantry

World War I with folding bicycles strapped to their backs

5.2.26

1899-

The Dursley-

A more

1917

Pedersen Cycle

comfortable

Company

woven net seat

designed and built and a

Fig. 5.2 (continued)

this innovative

lightweight

bicycle.

tubular frame.

Image

Land Transport

83

Figure

Date

Title

Innovation

5.2.27

1960s

Moulton ‘New

Small wheels,

Look’ standard

high pressure

M1 cycle built by

tyres,

Alex Moulton in

suspension on

Bradford on

front and rear

Avon, Wiltshire,

wheels.

UK. The Moulton bicycle was innovative, different and paved the way for modern folding bicycles. 5.2.28

1970s

Mountain

Suspension on

bicycles become

frame and fork,

popular –

large knobbly

designed for off-

tyres, heavy

road cycling.

duty wheels, disc brakes and lower gear ratios

5.2.29

1980s

Alexander

Aerodynamic

Vinokourov

design. Disc

competing in the

wheels work

London 2012

best where

Men's Olympic

there are no

Time Trial.

cross winds.

Track cyclists begin to use disc wheels or three, four or five spoke wheels.

Fig. 5.2 (continued)

Image

84

5 Moving Figure

Date

Title

Innovation

5.2.30

2005

A French Cyfac

New materials.

International Nerv Aluminium road racing bike

tubing and

built with Italian

carbon fibre

company,

parts with a

Campagnolo.

compact drive train.

5.2.31

2005

Giant Innova

Hybrid use for

hybrid city

touring and

bicycle with 27

mountain

gear speeds,

biking. Front fork and seat suspension, disk brakes for all weathers and heavy tyres to cope with urban hazards such as potholes

5.2.32

2008

2008 recumbent

Body weight is

bicycle built by

distributed over

Nazca Fuego,

a larger area

Nijeveen, The

with less

Netherlands.

pressure on the wrists, shoulders and seat. No saddle pain and lower air resistance.

Fig. 5.2 (continued)

Image

Land Transport

85

Figure

Date

Title

Innovation

5.2.33

2014

Italian Pinarello

Disc brakes,

Dogma 65.1.

aerodynamic

Seven of the nine

design and a

annual winners of

carbon fibre

the Tour de

frame.

Image

France to 2019 rode Pinarello bikes.

Fig. 5.2 (continued)

spokes. Some spokes were removable and could be replaced individually if they broke or were bent. People always want to go faster—the result was the penny-farthing. With its much larger front wheel a given turn of the pedals made the cycle travel further and hence faster. A cycle for two people—a tandem quadracycle—was built in 1886. A year later Henry Lawson made what he called the bicyclette with pedals turning a chain that drove a small rear wheel. Within another year performance in the USA was improved by using cup and cone ball bearings to reduce friction. British designers Hillman, Herbert and Cooper built the ‘Kangaroo’ in 1884—a bicycle with a gear and a chain drive on the front wheel. Mounting the bicycle was still a challenge so in the same year John McCammon of Belfast developed a bicycle with a step-through frame to make that easier. It was called a safety bicycle—safer than the penny-farthing. Also, that year the Humber safety bicycle was the first with a frame that we would recognise today. About this time tricycles became popular. Then in 1885 the Whippet safety bicycle had a type of derailleur gear with two speeds. The word derailleur seems to be derived from the French meaning the derailment of a train from its tracks. Some of the early designs used rods to shift the chain onto the various gears. In the same year Englishman John Starley called his safety bicycle Rover—a name that later became famous for motorcycles and cars. Scottish-born John Dunlop developed the pneumatic tyre in 1887 for bicycles then Frenchman Edouard Michelin produced a detachable tyre two years later. A safety bicycle just for ladies was produced two years after that in 1889. The 1893 Crypto cycle had a self-contained front hub gear and was the forerunner to modern 3-speed gear. By 1900 bicycles for gentlemen and bicycles for racing were being built. In 1920, the first bicycles for children were being made in an attempt to revitalise the industry. The first quick release hub came in 1930 and recumbent bicycles in 1934. The first BMX race was held in 1956 and the Moulton cycle appeared in 1962 and the mountain bike in 1970. The Japanese Tioga’s ‘Tension

86

5 Moving

Disk’ first produced in 1980s seemed to be a solid disc but actually was a tensionspoked wheel. The spokes were a continuous thread of Kevlar (aramid synthetic fibre) lacing the hub to the rim. The spokes were cased in a translucent disc for protection and better aerodynamics. In 1987, the first full suspension mountain bike was made by Paul Turner in Boulder Colorado and in 2013 Dutch cyclist Sebastian Bowler set a new world speed record for unassisted cycling in Nevada at 133.8 kph (83 mph) in his recumbent and aerodynamic velomobile. The first battery-powered bicycle was patented in 1895 but development was limited by battery technology. Modern e-bikes with pedal assist will develop over the forthcoming years though the legal distinction between them and mopeds and motorcycles could vary from country to country.

Engines Bicycles are engines that convert human energy into movement. Your body is a natural engine too in that it converts energy from the food you eat and the air you breathe into chemical energy in your muscles so you can move around and work. If you ride a horse, then the animal is your engine. If you freewheel down a hill on a bicycle you don’t need any muscle power, you just sit there and enjoy the ride. Instead, you are using gravity—converting potential energy due to your mass combined with gravity into the kinetic energy of motion. It took a long time to see that heat energy could be used to useful effect. Hero of Alexandria (First century AD) had used heat to do work but only to make gadgets which everyone dismissed at the time as merely toys. Then in the eighteenth century the developments came thick and fast as the primitive form of steam engine triggered some of the most remarkable ingenuity in the history of engineering. From that small but significant beginning we have created advanced heat engines—steam engines to power the railways, internal combustion engines to power cars—and most ingeniously of all jet turbines to power aeroplanes. Even the electricity you use every day mostly comes from a heat engine. Before we can look at how heat engines work, however, let’s just review some of the differences between fluids and solids.

Fluids and Solids Because fluids take on the shape of whatever is containing them, engineers cannot analyse the behaviour of large volumes of fluid in the same way as large chunks of solid. You will recall (Chap. 4) that solids have potential energy locked into them as they deform—engineers call it strain energy. They also have kinetic energy when they move. We said that, for solid materials, engineers must create structures that

Fluids and Solids

87

satisfy three necessary conditions, equilibrium, constitutive relations and compatibility. When a solid structure moves, say an aeroplane, ship or car, the separate parts stay together and move together—they obviously do not fly apart or accelerate away from each other. So, the parts of a moving structure remain connected by internal forces in a state of dynamic equilibrium. For fluids, engineers use a slightly different but equivalent approach. They don’t use force directly to check for equilibrium. Rather they use pressure which is the force on a small area of fluid. Likewise, instead of relying on mass they use density or mass per unit volume because the mass may move through a changing shape of the volume. Mass cannot be created nor destroyed—a principle called the conservation of mass. Engineers do use velocity for both solids and fluids—but for fluids they normally look at the average velocity of the particles at a cross section of, say, the flow in a pipe or channel. The quantity of flow is the average velocity times the cross-sectional area. Flow rate is flow per unit time. So, the equivalent of equilibrium, constitutive relations and compatibility for solids needs some thought for fluids. A useful way of expressing Newton’s Second Law for fluids is that force is the rate of change of mass multiplied by velocity. Mass multiplied by velocity is called momentum. So, another way of capturing the dynamic equilibrium of internal forces is that momentum remains constant. This is called the principle of the conservation of momentum. The reason this different but equivalent formulation is used for fluids is because, unlike solids, the flow of the mass has also to be considered explicitly. Then, just as for solids we have to define some constitutive equations. But now for fluids they relate mass, velocity, time, pressure, volume and temperature with density and other parameters. Also, just as for solids, we need compatibility between one state and the next. In fluids, this is usually referred to as continuity—of mass, energy or momentum—i.e. how those quantities are conveyed in space and time. For example, we may have continuous flow with no gaps or overlaps. Some fluids like blood, latex, honey and lubricants have much more complicated behaviour than air and water—they are called non-Newtonian fluids. Here we talk only of Newtonian so-called incompressible fluids. The language used by scientists is a bit confusing because compressible flow actually refers to situations where changes in pressure or temperature cause only negligible changes in density. As a consequence, at low speeds gases like air can be regarded as incompressible even though they do actually compress. At high speeds, however, the compressibility and the consequent changes in density have to be taken into account—for example, when analysing the airflow around a modern aeroplane. An example of simple but effective use of fluid flow is a Venturi meter—simply a constriction of the flow in a pipe. Engineers use it to measure flow rates of fluids. The idea came from Giovanni Battista Venturi who was a polymath of the eighteenth century. He was a priest, teacher of logic, a professor of geometry and philosophy and later mathematician, state engineer, for the Duke of Modena and responsible for bridges, dams, watercourses, and draining marshes. His discovered the ‘Venturi effect’, i.e. that the pressure reduces when water flows through a constriction in a pipe. The effect was turned into a practical metre almost one hundred years later by Clemens Herschel, who called his device a Venturi meter. Herschel was an American

88

5 Moving

bridge engineer who turned to hydraulic engineering through the influence of James Francis (who we meet later as the developer of the Francis turbine). Herschel was working for Holyoke Water Power Company in Massachusetts at the time and wanted to measure the amount of water used by individual water mills. Later, he worked on the hydroelectric power plant at Niagara Falls. Herschel’s innovation was to measure the pressure difference between the un-constricted and constricted sections of the pipe to calculate the flow in the pipe. The Venturi effect is not obvious—why should the pressure in a fluid change with the diameter of the pipe through which it is flowing? Imagine a liquid, like water, moving along a pipe at a steady rate or constant velocity before meeting a restriction. Then imagine what happens as the water reaches a constriction—it is rather interesting. The reason the pressure reduces is that the flow rate remains the same. So, to get the same amount of water through the smaller pipe then the velocity of the flow has to increase. Equivalently, when the diameter increases back again, the velocity will slow down again. In mathematical terms the flow Q = V × A or velocity V times area A. So, if flow Q remains the same and area A reduces then velocity V must get bigger. Pressure and velocity change with each other. The energy due to the velocity is kinetic. The energy due to the pressure is potential and the equivalent of strain energy in solids. You can see the release of potential energy, for example, from a propellant or pressurised gas if you set off a fire extinguisher. We find the Venturi effect in all sorts of places—in your blood vessels, arteries and veins, when plaque restricts the blood flow leading to the risk of a heart attack, in the carburettor of your car to measure airflow and ensure the correct amount of fuel is fed into the engine, in atomisers dispersing perfume or spray paint and in musical instruments with any narrowing of the wind pipe such as the mouth pipe of a trombone. The principle behind all of this is that of the conservation of energy. This says that energy can neither be created nor destroyed—but it can be changed from one form to another. In our case this is how pressure and velocity are related; if one changes the other must change in such a way that the total energy remains constant. In a Venturi when the velocity goes up then the pressure must go down to maintain the same total energy level. The problem for most of us is that common sense applied to a Venturi tells us that if you want to force a fluid through a smaller pipe you need to push harder, i.e. apply more pressure. That is because for a solid that is what you would have to do—force it through. That is no good in a fluid because a fluid is different from a solid in one essential way—it cannot resist shearing (sliding) forces between molecules—they just slip over each other. Therefore, the fluid cannot resist being pushed forward by reacting with increased pressure back—it can only respond by going forward faster. And the increase in velocity causes the pressure to drop.

Heat Engines Heat engines depend on fluid flow. The development of the heat engine was a turning point in the history of travel and moving. It is a story of ingenuity that led to cars,

Heat Engines

89

buses, trains and aeroplanes that are in use every day all around the world. But it has also brought pollution, congestion and is using up the world’s supply of nonrenewable fossil fuels—things will have to change. In the seventeenth century, Samuel Morland was one of the first to wonder if heat could be used more advantageously. He experimented with gunpowder to make a vacuum, so water would be sucked up, to form a water pump. There was a very practical need—mines were often flooded. That prompted Thomas Savery in 1702 to build his ‘Miner’s Friend’—an engine to raise water by the ‘impellent force of fire’. This first steam engine had no piston. He heated water in a boiler using fire in a furnace. He had a closed container with three connecting pipes—each with valves. The first steam pipe connected with the boiler. The second suction pipe connected with the mine water, and the third outlet pipe connected to the outside. The cycle started with the steam pipe valve open and the other two pipe valves closed so the cycle was as follows. First, the steam from the boiler passed into the container. When the container was full of steam, the steam valve was closed so that the container was entirely isolated. Second, the container was sprayed with cold water so that the steam condensed and created a vacuum. Third, the valve on the second suction pipe was opened and consequently the mine water was sucked up the pipe to fill the vacuum. Fourth, the suction pipe valve was closed, and the outlet valve opened. Finally, the steam valve in the pipe from the boiler was opened again and the steam drove the mine water out of the container through the outlet pipe to waste. The system could not operate if the mine water was deeper than 10.3 m because that is the limit for a vacuum removing atmospheric pressure. Savery compared the work of his engine with that of horses and so the term ‘horsepower’ was born. Thomas Newcomen used the same idea as Savery, a vacuum from condensed steam, but he used it to pull a piston down a cylinder rather to suck up mine water. The piston was linked to the end of a timber beam which rotated about a central pin. The other slightly heavier end of the beam was connected to a pump within the mine water. As the piston rose and fell, the beam rotated and the pump rose and fell—bringing water out of the mine (Fig. 5.3). In 1765, James Watt improved Newcomen’s engine by incorporating a separate condenser and then returning the warm condensed water to the boiler. As nineteenth century opened, the railway age was about to begin. Richard Trevithick built an engine with much higher steam pressures (30–50 psi or 0.2– 0.35 N/mm2 ) and used it to power a locomotive that ran 9 miles from ironworks at Penydarren to the Merthyr-Cardiff canal in South Wales. From that moment on, moving around was set to change dramatically. The railway network in the UK flourished between 1830 and 1870. Almost every significant centre of population in England and Scotland was connected to the railway. The landscape was transformed both physically and culturally. Newspapers, fresh produce such as milk and meat and manufactured goods were distributed daily. People could travel for holidays. Letters could be exchanged often on the same day. New opportunities for business opened up. An express train could travel at 80 mph. By 1863 traffic in London was becoming so congested that the Victorians moved some of it underground. The first

90

5 Moving

Fig. 5.3 Beam engine

underground railway connected Paddington station to Farringdon Street about a mile from the Bank of England. Steam engines were the first heat engines. Heat engines create useful work from a difference in temperature. An energy source (e.g. burning fuel—chemical or nuclear) creates heat which does mechanical work. The science of understanding and designing heat engines is called thermodynamics. The law of conservation of energy I mentioned earlier is also called the first law of thermodynamics. That is because the change in the internal energy of a closed thermodynamic system is equal to the amount of heat supplied to the system minus the amount of work done by the system on its surroundings. A closed system is one with walls through which matter cannot pass but energy can—as heat or work. Engines have to provide continuous power, so they work through a series of thermodynamic processes that cycle round and round—thermodynamic cycles. In theory, the processes are reversible. In practice, they are irreversible as some energy gets lost and unavailable to do any work. An important measure of that loss is called entropy. Entropy is a difficult concept, but it is probably best thought of as an increase in disorganised energy at the expense of organised energy. Sadi Carnot was a French military engineer and is generally acknowledged as the ‘father of thermodynamics’. He asked himself whether there was a limit to the number of improvements that could be made to a steam engine. He looked closely at the way

Heat Engines

91

they operated and realised that the cyclic process begins and ends with water. Heat is added to water to make steam. The steam expands and does mechanical work by pushing a piston. Finally, the steam is condensed back into water. He reasoned that if the inefficiencies arise only from leakages and friction then he could imagine an engine without them—an ideal engine with a perfectly insulated cylinder and a leakproof frictionless piston. He realised that because the steam had to be condensed to water then some heat loss was inevitable—even in this ideal situation. We know now that Carnot’s ideal heat reversible cycle can’t be bettered because there are always losses in any practical engine—but we can use it as a comparator. Carnot’s theory did not have any significant immediate practical effect, but it did demonstrate that a heat engine is more efficient the higher the temperature difference within the cycle. In 1850, German Rudolph Clausius formulated the first law of thermodynamics and in 1854 the second law of thermodynamics. The second law captures the common-sense notion that you can’t get something for nothing. As we have already seen, some of the energy is dissipated or becomes irretrievable and no longer available to do useful work—Clausius was the first to call this entropy. He said that it is impossible to cause heat to flow from a cold to warmer body unless we supply extra energy. He was the first to show that no engine could be more efficient than the reversible Carnot cycle. The piston steam engine we have looked at so far is an external combustion heat engine since the heat source is external to the cylinder. Cars and trucks have internal combustion engines (IC) because the fuel (hydrocarbons such as petrol—USA gas— and diesel) combusts inside the engine. The two main types of IC are spark-ignition engines and diesel engines. Sparkignition engines work on a thermodynamic cycle—the Otto cycle—named after its German engineer inventor Nicolaus August Otto who took a patent in 1876. Inside an internal combustion engine are one or more hollow cylinders (Fig. 5.4). Each one is closed at one end and open at the other. Inside each cylinder is a close-fitting piston that moves up and down—the distance it moves is called the stroke. The engine goes through thermodynamic cycles—some engines have four strokes per cycle and some have two strokes per cycle. In a four-stroke engine, the air–fuel mixture is sucked into the cylinder through an intake valve as the piston moves down the cylinder—this is the first or intake stroke. As the piston moves back up, driven by a flywheel as we shall see, the intake valve closes, and the mixture is compressed—this is the second or compression stroke. When the piston is near to the top of the cylinder (called top dead centre TDC), the fuel is ignited usually by a spark. The burning mixture drives the piston downwards—the third or power stroke. As it does the up-down reciprocating motion is converted to rotary movement by a crankshaft. Then via a system of gears, the shaft drives the wheels of a vehicle. When the piston reaches the lowest point, again the exhaust valve opens, and the gases are expelled as the piston moves back up—the fourth or exhaust stroke. The cycle then repeats as the cylinder moves down through the next intake stroke. The reason it can repeat is because a flywheel mounted on the rotating shaft has developed enough momentum to drive the piston back upwards. The nature of the four-stroke cycle means that the more the cylinders there are in the engine, the smoother it will be. There are different ways of orienting the cylinders

92

5 Moving

Fig. 5.4 The Otto cycle

relative to each other—common ones are simply in-line whilst others are arranged in a V shape. The firing of each cylinder has to be carefully ordered so that torsion on the shaft is even and the engine balanced. For four in-line cylinders numbered 1–4 typical firing orders are 1-3-4-2 or 1-2-4-3. For a V-shaped engine, it might be 1-4-2-3. Engines are only part of the engineering of cars. Aerodynamics is also important for fuel economy and speed—none more so that in motorsports. A split second can be the difference between winning and losing and worth a great deal of money. In a racing car, man and machine are one—they must be to get maximum performance. Professor Caroline Hargrove specialises in making that happen. She has been the Technical Director at McLaren Applied Technologies since 2013. She pioneered the simulation of interactions between human and machine, first in motor racing and then in sport and medicine. She obtained her Ph.D. in Engineering and was a lecturer at Cambridge University before moving into industry. She works with top racing drivers and so travels around the world with them. She says that proving your worth in a competitive, male-dominated environment is challenging. She leads the engineering teams looking after motorsport, elite and professional sports, transport, energy and medical services. She is also a visiting professor in the Department of Surgery at Oxford helping to assess surgical performance and improving care. She says that multidisciplinary teams and culturally diverse environments are essential for companies to succeed. She thinks that inspiring girls to study STEM subjects is the responsibility of all female engineers today, and by creating more female-friendly workplaces, companies can become more innovative and successful. At the beginning of the twentieth century, cars were a toy for the rich. Now cars dominate our lives as they moved from luxury to essential—and now to threat—the

Heat Engines

93

principle of the Unintended again. An automotive trade journal has estimated that the total number of motor vehicles (cars, trucks and buses) in the whole world crossed 1 billion in 2010 and if present trends continue there will be 2 billion by 2035.4 The consequences are serious. Traffic is clogging up our cities, polluting the air and through carbon emissions contributing to climate change. But there is no going back. We cannot undo what has been done so we have no choice but to move forward by evermore ingenious engineering. Driverless electric cars are one way in which the future will be quite different to the past. Victor Wouk was the American engineer who has been dubbed the grandfather of electric and hybrid vehicles in the United States. He graduated from the California Institute of Technology and worked for Westinghouse in Pittsburgh, where he developed high-voltage controls for centrifuges used to purify uranium for the Manhattan Project. In the 1940s and 50s, he founded and sold two electric manufacturing and supply companies. In 1962, one of the founders of Motorola asked him to explore the possibilities of an electric car. Wouk knew that batteries were the key, so he started to design a system that would combine an internal combustion engine with an electric motor. Wouk’s idea was approved by the United States Environmental Protection Agency (EPA) ‘Clean Car Incentive Program’ in 1971. EPA was ‘to consider a nationwide test of vehicles based on his design if satisfied with the prototype’. Wouk and friends invested about $300,000 and successfully converted a Buick sedan—the first full-sized hybrid vehicle featuring a 20-kilowatt direct-current electric motor and a Mazda rotary engine. Fuel economy was doubled with emission rates around 9% of those of an IC engine of that time. But the design was rejected—the reason given is that bureaucrats decided there was no future in the idea and Wouk ran out of money. In 2000 and in his 80s, Wouk bought a white Toyota Prius—the first hybrid car. Since 2008 due to advances in battery design, increasing oil prices and concerns about pollution electric vehicles are enjoying something of a renaissance. Unfortunately, hybrid cars alone will not reduce greenhouse gas emissions sufficiently and all-electric cars powered from coal or oil sources will still be serious polluters. However, all-electric cars powered from renewable or nuclear energy will be low- or zero-carbon emitters. It is quite possible that soon we will consider internal combustion engines to be ‘old technology’. The new ‘engines’ will be batteries or fuel cells separately or combined. Fuel cells are much like batteries but instead of being charged from the electrical mains they are fuelled by hydrogen. The cell produces electricity through a chemical reaction, without combustion, that converts hydrogen and oxygen into water. Batteries provide high levels of power (rate of using energy) but do not store energy well. Fuel cells provide much more energy but their power per unit weight is limited. Cars need short bursts of high power to start and accelerate so low capacity batteries could be used with fuel cells for steady running. Efficiencies of electrochemical ‘engines’ compared to the internal combustion engine are much higher. Electricity can also be produced from light using solar cells. Solar-powered aeroplanes are coming. In 2016, Bertrand Piccard and André Borschberg completed the first circumnavigation of the globe in one. The Solar Impulse prototype plane (Fig. 5.5) has the wingspan of a jumbo jet, the weight of a family car, and the power

94

5 Moving

Fig. 5.5 Solar Impulse aeroplane. Image taken from blog by David MacKay

of a small motorcycle. The solar panels collect the sun’s rays and convert them into electricity. The electricity is then either sent directly to the motors or, if the energy generated is greater than the energy needed to fly, it is sent to the batteries. During the day, the plane flies only by the energy from the sun. But in the morning and evening, when sunshine is not so strong, and especially at night, the plane taps into the batteries. Every evening, the pilot must make sure that the plane’s batteries are 100% charged so that it can fly until the next sunrise. The batteries have an energy density (the amount of energy per unit weight) like a car battery and in total were just over a quarter of the aircraft’s all-up weight.

Turbines What do a waterwheel, windmill, power station and jumbo jet have in common? They all use or rely on turbines. The waterwheel and the windmill are turbines but not heat engines and the power station and jumbo jet rely on turbines that are heat engines. Turbines are ingenious engines that are very important for modern life. Turbines work differently from the engines we have looked at so far in that they capture energy from a moving fluid—water, wind, steam or heated gas. The word turbine derives from the Latin word for whirling or vortex. All turbines work in essentially the same way—the fluid turns the blades of a wheel mounted on a shaft. A set of gears transmit the rotation to the shaft of a grinding mill, the axle of an electrical generator or the wheels of a vehicle. The simplest turbine is a child’s windmill or pinwheel—you blow to make it spin. If the toy is robust enough, you could hold it under the tap and the water would make

Turbines

95

it spin too. Any flowing fluid will do the job. When you hold your hand under a cold water tap in the bathroom you feel the flow. As you turn up the tap, the force on your hands gets bigger. Note that when you turn the tap you do not increase the water pressure rather you increase the flow. The different types of turbine evolved interdependently but are engineered differently. A key factor in the engineering is whether the fluid is compressible or not. That is because the forces are transmitted through the fluid differently. The flow of an incompressible fluid is immediate. All fluids are compressible to some degree but some like water are effectively incompressible. For example, when you suck water up a drinking straw you get the water straightaway. The same is true when water flows along a long pipe. Air can be assumed to be incompressible at low velocities but not otherwise. The flow of a compressible fluid is not so immediate—rather a pressure pulse travels through the fluid. The higher the density of the fluid, the slower the speed. Turbines are also either impulse or reaction machines. Impulse turbines turn a wheel by the impact of a fluid on blades attached to a wheel. Consequently, the direction of the flow of the fluid is changed. The child’s windmill under the water tap is an impulse turbine because a fluid is impacting on a surface of the wheel and making it rotate. Reaction turbines work differently—they rely on the reaction to the pressure or force of a fluid. They turn a wheel as the fluid flows through the blades like a windmill—so the direction of the flow is unchanged. A common example is a garden sprinkler where water enters the arms of the sprinkler at low velocity and leaves through jets at high velocity. Practical turbines are examples of the various ways that the flow of water, wind, steam and gas can be manipulated—in other words we coax natural forces to do something we want. The ways in which waterwheels and windmills work are reasonably transparent—water and wind turning wheels. But, as we shall see, steam and gas turbines are ingenious and imaginative but opaque and hence difficult to see how they work. Generating electricity from waterpower is sometimes called hydrokinetics or hydropower. Water held high in the mountains or low level in the tides flows naturally. Water stored in a tank or reservoir can be released in a controlled way. In both cases, the water, held at height, has potential energy—often called pressure head—then as the water flows the potential energy converts to kinetic energy which the turbine converts to a rotary motion that drives an electrical generator. Since the kinetic energy is due to the velocity of the flow of water, it is often called velocity head. The ancients used water wheels—of course not to generate electricity—but to mill and grind corn. They were wooden or metal wheels with blades or buckets attached on the outer rim. Water hit the blades and made the wheels rotate. Gear mechanisms turned shafts to power the milling. You see water wheels even today—but they are rarely used commercially. Water wheels could be vertical or horizontal. They are called undershot when the water strikes the wheel at the bottom, breast shot in the middle and overshot at the top. They were usually fed from a millpond formed by

96

5 Moving

damming a river or stream. The channel for taking the water to and from the wheel is called a mill race. Lester Pelton invented the Pelton wheel. Lester was born in Ohio but attracted to the California gold rush in 1850. Samuel Knight had built an impulse water wheel using a high-pressure nozzle to focus water onto cups mounted on a wheel. Lester noticed that when the jet on a water wheel hit the cups near the edge, rather surprisingly, the wheel moved faster. He discovered that when the jet hit the middle the splash wasted energy. His redesign rather resembles two cupped hands making a split bucket. He patented it in 1888, and the two wheels were in direct competition for a while. Pelton wheels are still used when the water has a high hydraulic head (i.e. height above a datum point) or pressure with a low flow rate. You may see an another different example of an impulse turbine when you visit your dentist. Many dentists use a high-speed drill that has air flowing through multiple nozzles onto turbine blades which rotate and turn the drill. In the USA in 1837 an Englishman, James Francis, was appointed Chief Engineer to the Locks and Canal Company on the Merrimack River, Lowell, Massachusetts. With a colleague, Uriah Boyden, he developed what is now one of the most widely used reaction water turbines in the world. The Francis turbine is used in almost all the hydroelectricity dams built since 1900 and generates perhaps a fifth of all the electricity of the world—including the massive Three Gorges Dam in China. A Francis turbine (Fig. 5.6) is an inward flow turbine because the fluid enters at the outer periphery and flows inwards towards the centre and discharges at the outer periphery. It has four main parts—a spiral casing, guide vanes, runner blades and a draft tube. The spiral casing is also known as the volute. It has openings at regular intervals to allow the water to impinge on the runner blades. The job of the guide or stay vanes is to guide the flow and convert the potential energy due to pressure into

Fig. 5.6 The Francis turbine inlet scroll Grand Coulee Dam. Right image: Image by the US Bureau of Reclamation, Public domain via Wikipedia Commons

Turbines

97

kinetic energy. The water strikes the runner blades and the shaft rotates. The angles of the blades are important. A so-called draft tube takes the water from the runner blades and discharges it. Viktor Kaplan in 1912 used the Francis turbine as a basis for a new inward flow reaction turbine with axial flow, i.e. flow parallel to the shaft. The major change was to use adjustable blades (Fig. 5.7) and inflow gates so that the turbine can cope efficiently with varying flow. In that way, Kaplan was able to improve efficiency over a range of flow rates. The Kaplan turbine is widely used today where pressure heads are low and flow rates are high. Windmills obviously catch the kinetic energy of the wind. The faster the wind blows, the more energy it contains and the faster the sails spin. Wind turbines generating electricity turn relatively slowly, largely for safety, so they are big to capture enough energy. You can’t tell by looking—but the wind blows slightly more slowly after it’s passed by a windmill. The blades or sails are a key part of the design and carefully shaped to catch most energy. Each blade works like an aeroplane wing—as air passes around both sides of the blade the shape creates higher air pressure on one side than the other—and so it spins. Steam is used in many power stations to generate electricity. Steam turbines are heat engines that were pioneered by British engineer Charles Parsons around 1884. Water is heated by burning coal or oil (or in a nuclear power station through a nuclear reaction) to make the steam which is pushed through a turbine. Water and wind turbines have only one rotating set of blades. Parsons realised that the velocity of the stream flowing over the blades had to be kept down to avoid damage. His ingenious idea was to use a number of turbines in series (known as stages) and arranged in a sequence inside what is effectively a closed pipe. The steam enters the pipe and is channelled past each stage in turn so progressively more of the energy is extracted. If you’ve ever watched a kettle boiling, you’ll know that steam expands and moves very quickly if it’s directed through a nozzle. For that reason, steam turbines turn at very high speeds—many times faster than wind or water turbines. Gas turbines rely on gas, rather than steam, going through a Brayton cycle. In 1872, American George Brayton applied for a patent for an internal combustion reciprocating piston engine with a constant gas pressure. It was based on a very early suggestion by Englishman John Barber in 1791 who tried to design a horseless Fig. 5.7 Diagram of Kaplan turbine

98

5 Moving

carriage. You will recall that a thermodynamic cycle is a linked sequence of processes in which heat and work are transferred in and out of a system as pressure, temperature and volume changes. The Brayton cycle has four stages. First, air is drawn in. Second, it is compressed so the pressure rises. Third, the air goes into a combustion chamber where it is mixed with fuel and ignited so the temperature rises but the pressure remains the same since the chamber is open. Fourth, the heated, pressurised gas passes through a series of turbine blades and is ejected from the rear. Modern gas turbines are made from especially resilient alloys because they work at such high temperatures. Jet aeroplane engines work essentially this way too—summarised succinctly as ‘suck, squeeze, bang and blow’. Air is sucked into the front of the engine, compressed, mixed with fuel and ignited and blown out at the back. But in 1920 William Joseph Stern reported to the Royal Air Force that there was no future for the turbine engine in aircraft. He thought that the existing designs of compressors had too low efficiency. Frank Whittle was set to change that.

Jet Engines Frank Whittle joined the RAF in 1923 and after early setbacks was admitted to RAF Cranwell—the training school in Lincolnshire for RAF officers. His final year thesis was about motor jets or conventional piston engines providing compressed air to a combustion chamber and then using the exhaust gas to provide extra thrust—a kind of after burner to a propeller engine (An afterburner is a component added to some jet engines to increase thrust by injecting more fuel into the jet stream after the turbine to reheat the gas, increase the pressure in the tailpipe and eject the gas at a higher velocity). Whittle, by all accounts, was an adventurous pilot. He was disqualified from an end of term flying contest because of his dangerous flying. He continued to work on his motor jet and had the idea of replacing the piston engine with a turbine powered by the exhaust to drive a compressor. He showed his idea to his commanding officers who encouraged him to send it to the Air Ministry. The Ministry consulted Griffith who pointed out an error in Whittle’s calculations and decided that the idea was impractical. Whittle patented his proposal in 1930. In 1932 he graduated from the Officer’s Engineering Course at RAF Henlow in Buckinghamshire with exceptional marks. So exceptional he was permitted to take a two-year course at the University of Cambridge from where he graduated with a First in Mechanical Sciences in 1936. In 1935 Whittle’s patent lapsed. After some negotiations, a partnership with financial backing was formed in 1936 to create Power Jets Ltd. Similar work was being done in Germany. In 1930, Paul Schmidt had patented a pulse jet engine which later became the buzz bomb or doodlebug or the V-1 flying bomb. A pulse jet is relatively simple. Fuel and air are ignited at the intake, the intake valve closes and the combustion gases travel down a pipe leaving at the rear and creating a thrust. As this happens the pressure behind the moving gas drops so the intake valve opens, and more air and fuel drawn in from outside. But at the same time the low pressure causes some of the hot combustion gas to be drawn back towards the intake. The pressure rises, and the intake valve closes and when the two meet the cycle starts again. This pulsating

Jet Engines

99

motion can create a buzzing noise. In 1935, Hans von Ohain had plans for a turbojet engine and convinced Ernst Heinkel to develop a working model and together they developed the first turbojet on a test stand. In 1937, the Heinkel HeS 1 experimental hydrogen-fuelled centrifugal jet engine was tested and in 1939 the plane flew for the first time. Despite that, the priorities of the UK government during World War II were elsewhere. Whittle had great difficulty in convincing the government that his ideas were worth investment. By the end of 1936, a prototype was well underway and the following year the first engine ran successfully. Further delays in funding hindered the work but in 1939 the Air Ministry decided to help by injecting some cash and placing orders for flying versions of the engine. By 1941 this was achieved and for this Whittle is often called the ‘father of the jet engine’. The Air Ministry issued contracts to the Rover Company to build engines but the collaboration between them and Power Jets ran into difficulties. Whittle was frustrated by Rover’s inability to meet the required standards. Rover set up their own operation and this caused a major crisis. In 1940, Whittle met Stanley Hooker of Rolls Royce who introduced him to Ernest (later Lord) Hives. Rolls Royce offered to help develop Whittle’s engine by building a test rig. In 1942 they were asked to build six engines. Roll Royce and Rover did a deal in 1942, the Rover plant was closed, and the first production engine was the Rolls Royce Welland in 1943 for the Gloster Meteor—soon to be replaced by the RR Derwent in 1944. The continuous stress on Whittle resulted in nervous breakdowns in 1940 and 1944. He retired from the RAF and was knighted. He immigrated to the USA in 1976 and died of lung cancer in 1996. Recall that a gas turbine jet engine has the same four thermodynamic stages as the internal combustion piston engine in your car. But it is much more elegant and most ingenious because rather than happening intermittently, the stages occur continuously and are mounted on a single shaft. The pressure and temperature in a piston engine change quite dramatically with time whilst in a turbine both the pressure and temperature remain constant at steady speeds at given locations in the engine. The gas turbine is a very clever way of manipulating the pressure, volume, velocity and temperature of gas to create the thrust that propels the aircraft. It is another example of aesthetic of function—an elegant flow of energy. A turbine may not be beautiful form but the way it works is of great sophistication and artistry. The engine is a working example of Newton’s third law of motion—that for every action there is an equal and opposite reaction. It’s rather like the trick you will probably have done at some time with a toy balloon. You blow it up and release it so that the air rushes out creating the thrust that makes it fly away. In an engine, the air is taken in at the front and expelled at the back. The size of the thrust depends on Newton’s second law of motion which, you will recall, states that a force (thrust) is equal to the rate of change of momentum or mass multiplied by velocity. In other words, the thrust depends on the mass of the flow of air through the engine and the difference between the velocities of the air entering at the front (the speed of the aeroplane itself) and leaving at the rear. An axial compressor in a gas turbine looks rather like a fan but, as mentioned earlier, has a set of specially shaped rotating radial blades, called rotors, mounted

100

5 Moving

on a disc on the central shaft. Alongside each rotor is another set of blades, again specially but differently shaped, called stators but these are fixed in position and do not rotate. As the air passes through each set of rotors its velocity increases and, as it passes through the stators, the gas is diffused turning this kinetic energy into pressure energy—hence, the velocity fluctuates but remains essentially the same whilst the total volume of the gas reduces, and the pressure ratchets up by a factor maybe of the order of 20–40 times. The temperature also increases to maybe 500 °C but the more efficient the compressors, the less the temperature rise. The high-pressure gas then passes along to the next stage—combustion. Here the fuel (propane, natural gas, kerosene or jet fuel) is injected through a ring and burnt. The temperature rises dramatically but the pressure remains essentially the same as the volume increases. The hot high-pressure gas at maybe around 1,600 °C is then accelerated into the turbine by reducing the volume. The turbine has blades like the compressor but shaped differently. The gas is guided by vanes or stators through the rotors expanding as it does and spinning them (essentially a compressor in reverse). The remaining high-pressure gas is then expanded to rush out of the exhaust at high velocity to produce thrust (similar to the balloon). The materials used in the turbine melt at around 1,200 °C, so they must be cooled. This cooling technology applied to a blade made of ice would keep that blade frozen even in the hottest domestic oven. The compressor is mounted on the same shaft as the turbine and so the turbine spins the compressor at speeds of around 3,000–10,000 revolutions per minute at take-off. The forces on the blades as they spin at these high speeds are considerable and so they have to be specially designed to stop them breaking up. Modern jet engines in commercial passenger aeroplanes are often turbofan jets (Fig. 5.8). These are gas turbines with a large fan at the front to suck in more air. The thrust of the engine then comes from two sources: the first is the gas turbine itself and the second is ‘bypass air’—so called because it bypasses the turbine portion of the engine and moves straight through to the back of the nacelle (the engine housing) at high speed. The fan may be very big—of the order of 3 m in diameter so it can move a lot of bypass air and hence creates much more thrust very efficiently. The speed of the bypass exhaust air is less than that from the turbine and so the average speed is lower. Since engine noise depends on the speed of the exhaust gases, the turbofan jet engine is quieter.

Moving On In summary, in this chapter, we have seen that moving requires energy—so the engineering of moving is about managing the way energy flows. We coach natural forces to do what we want. All processes have energy as a potential drives a flow. It took a long time to the initial idea of a bicycle but then the evolutionary progress was relatively quick. The ideas of a few individuals were influential in triggering yet more ideas into an explosion of innovation. Likewise, it took us a long time to understand that we can use heat to power machines—then again relatively quickly along came the steam engine and out of that, the internal combustion engine, the

Moving On

101

Fig. 5.8 Turbofan Jet engine

turbojet and the space rocket. Although the steam engine was at first an intuitive notion, later developments depended on an understanding of the essential technical differences between solids and fluids. Our relatively recent discoveries of how to generate and use electricity are becoming even more important as we enter the age of the electric vehicle. The future of travel will change—at once exciting and portentous. Congestion, pollution and our depletion of non-renewable fuels have become major challenges. IT is already having an effect. We have apps to summon taxis and share cars. We have cars that will park themselves. ‘Smart cities’ are beginning to use many types of electronic data collection sensors to supply information (for example, levels of pollution) that can be used to manage assets and resources. We are forecast soon to have driverless cars, trucks and buses and trains. We will have high-speed trains possibly through hyperloops—sealed tubes with reduced air resistance.5 These are all developments that will reduce congestion through personal transport that will travel faster and closer reducing wasted time. Even bigger cruise ships will cruise our oceans. However, until we reduce our dependence on fossil fuels, we are only redistributing the big problems—carbon emissions and depletion of finite resources. We can sequester some of the carbon but the best way to do it remains a challenge. Renewable energy is key to making electric vehicles non-polluters. Solar energy equipment will be improved and become cheaper. Roads, buildings and other spaces will have solar panels built into the surface. Wind energy will continue to be developed. Information technology will also reduce the need to travel as we shop online and arrange virtual meetings. Videoconferencing may well involve life-size 3D holograms. Rockets and new spacecraft will widen the possibilities for new ventures in space travel.

102

5 Moving

The key to all of this unfolding is control and communication—both the physical systems of electromagnetism (i.e. electricity and magnetism) and the people and social systems in which they are embedded. We will need to keep in mind our five principles as the Unintended consequences hit us—particularly those of Artificial Intelligence and robotics. The key is digital technology so that is what we turn to in the next two chapters—the ingenious but increasingly opaque age of information technology and complex systems. How will we ‘learn our way’ through the consequences of such potential massive change? End Notes 1. Bijker, WE., (1997) Of bicycles, bakelites and bulbs. The MIT Press, Cambridge, Massachusetts, USA. 2. For quotes by Albert Einstein and by Susan B. Anthony see http://theargonauts. com/bicycle-quotes/ (Last accessed February 2019). 3. Figure 5.2 Credits Figure 5.2.1: Image by John O’Neill. CC BY-SA-1.0 via Wikipedia Commons. Figure 5.2.2: Image by Bellicorum Instrumentorum Liber. CC BY-SA2.5 via Creative Commons. Figure 5.2.3: Image by Gadfium. Public domain via Wikimedia Commons. Figure 5.2.4: Image by Gun Powder Ma. CC BY-SA-3.0 via Wikipedia Commons. Figure 5.2.5: Image by Ian Wilkes copied to commons by Jarry 1250. CC BYSA-3.0 via Wikipedia Commons. Figure 5.2.6: Image by Ian Wilkes copied to commons by Jarry 1250. CC BYSA-3.0 via Wikipedia Commons. Figure 5.2.7: Image taken from alchetron.com, Kirkpatrick Macmillan Wikipedia. CC BY-SA 3.0. Figure 5.2.8: Copyright permission granted by Colin Kirsch at https://oldbike. eu/1880s-tiller-treadle-adult-tricycle/. Figure 5.2.9: Image in Public domain via Wikipedia Commons. Figure 5.2.10: Public domain via Wikipedia Commons. Figure 5.2.11: From J.T. Goddard, ‘The velocipede: its history, varieties and practice’. Image in Public domain via Wikipedia Commons. Figure 5.2.12: From J.T. Goddard, ‘The velocipede: its history, varieties and practice‘. Image by Agnieszka Kwiecie´n, CC BY-SA-3.0 via Wikipedia Commons. Figure 5.2.13: Image uploaded by Darwinek and in the Public domain via Wikipedia Commons. Figure 5.2.14: Image by Manotechnologie. CC BY-SA-3.0 via Wikipedia Commons. Figure 5.2.15: Image by Ian Wilkes, CC BY-SA-3.0 via Wikipedia Commons. Figure 5.2.16: Image by Ian Wilkes and copied by Jarry 1250. CC BY-SA-3.0 via Wikipedia Commons. Figure 5.2.17: Image by Ian Wilkes and copied by Jarry 1250. CC BY-SA-3.0 via Wikipedia Commons.

Moving On

103

Figure 5.2.18: Image by Motty. CC BY-SA-3.0 via Wikipedia Commons. Figure 5.2.19: Image by Ian Wilkes and copied by Jarry 1250. CC BY-SA-3.0 via Wikipedia Commons. Figure 5.2.20: Image by Flowizm. CC BY-SA-2.0 Generic licence via Wikipedia Commons. Figure 5.2.21: Image from Bartleet’s Bicycle Book by Horace William Bartleet (1870–1943) published in 1931/2 by Ed. J Burrow & Co Ltd, London See Grace’s Guide to British Industrial History. Figure 5.2.22: Image from cover of the Petit Journal Extrait du Patrimoines brestois N°4 CC-BY-ND via Wikipedia Commons. Figure 5.2.23: Image originally from Bartleet’s Bicycle Book by Horace William Bartleet (1870–1943) published in 1931/2 by Ed. J Burrow & Co Ltd, London See Grace’s Guide to British Industrial History. Figure 5.2.24: Image in the Public domain. Transferred by Idleguy from en.wikipedia to Wikipedia Commons. Figure 5.2.25: Image in Public domain. Uploaded by Risorgimento at English Wikipedia and transferred from en.wikipedia to Commons. Original text: Lombardi Historical Collection. Figure 5.2.26: Image by Yesterdays Antique Motorcycles en Classic Motorcycle. CC BY-SA-2.5 via Wikipedia Commons. Figure 5.2.27: Image in Public domain and image uploaded by G-Man to Wikipedia Commons. Figure 5.2.28: Image by Andy Armstrong at the 2005 Dusk ‘til Dawn event in Thetford Forest, UK. CC BY-SA-2.5 via Wikipedia Commons. Figure 5.2.29: Image by David Iliff. License: CC-BY-SA 3.0 via Wikipedia Commons. Figure 5.2.30: Image by Julius Kusuma, from English Wikipedia. CC BY-SA-3.0 via Wikipedia Commons. Figure 5.2.31: Image by Ian Donaldson. Originally uploaded by Icd at English Wikipedia. CC BY-SA-3.0 via Wikipedia Commons. Transferred from en.wikipedia to Commons by Sreejithk2000. Figure 5.2.32: Image by P.E. Howland at English Wikipedia. Public domain via Wikipedia Commons. Transferred from en.wikipedia to Commons by Javier Carro. Figure 5.2.33: Original image by Glory Cycles, Greenville, SC, USA. CC-BY-2.0 at Wikipedia Commons. 4. See http://wardsauto.com/news-analysis/world-vehicle-population-tops-1billion-units (Last accessed February 2019). 5. A hyperloop is a sealed tunnel or tube through which a train or pod may travel with reduced air resistance to travel at high speed. See https://en.wikipedia.org/ wiki/Hyperloop (Last accessed February 2019). Musk, Elon (2013). Hyperloop Alpha. SpaceX. See http://www.spacex.com/sites/spacex/files/hyperloop_alpha20130812.pdf (Last accessed February 2019).

Chapter 6

Communicating

Nature is a tapestry of patterns

Patterns What does a weaving loom and a computer have in common? The answer is that one creates patterns of threads to make cloth and carpets—the other creates patterns of pulses of electromagnetic energy to make information. If that sounds all a bit bizarre or far-fetched—it isn’t. There was a direct connection. Charles Babbage the inventor of the Analytic Engine and the first programmable computer in 1837 was inspired by the loom developed by French weaver Joseph Marie Jacquard in 1804. The first question for this chapter is this: why are patterns so important in human communication? Secondly, how influential is the distinction between the syntax or structure of patterns and their semantics—in other words, how we interpret them to give them meaning? Behind these questions, we should remember that we would have no patterns of electromagnetic energy were it not for the foundational work of Michael Faraday. Not only was he a key figure in the development of the electrical machines, as we witnessed in the last chapter, but his influence led to the work of one of the greatest physicists of all time James Clerk Maxwell and the subsequent development of the telephone, wireless radio and modern telecommunications equipment. How important in those developments was the invention of the ‘valve’, the transistor and the integrated circuit? The latter is present in all modern electronic equipment such as mobile/cell phones and computers and it is there we find the patterns of electromagnetic energy. How do digital electronics and computers work with those patterns? Should we be excited or worried about the ever-increasing power of the patterns of artificial intelligence?

© Springer Nature Switzerland AG 2020 D. Blockley, Creativity, Problem Solving, and Aesthetics in Engineering, https://doi.org/10.1007/978-3-030-38257-5_6

105

106

6 Communicating

Patterns are critical in human communication—in language, body movement, dress, ornamentation, music and dance. When we talk, write, behave or perform we send out signals—patterns of light, sound and touch which are received and interpreted by others. Patterns are everywhere—woven into cloth and carpets, letters written on a page, daubs of paint in a picture, pixels in a camera, bar codes for a laser scanner, sounds of the human voice or music, shapes carved into a sculpture, or connections between body cells such as neurons in the brain. From the precision of mathematical and logical deduction to the style and rhythm of poetic expression of feelings and emotions, patterns shape us. Patterns are also key to the engineering of communication. Importantly, and perhaps surprisingly, we find that patterns of holes punched into cards evolved into arrays of ‘bits’ of distinct levels of electrical potential (volts) held in the memories of computers and ubiquitous in all forms of modern communications engineering. Semiotics, the study of signs and symbols to communicate, is usually divided into three branches: syntax, semantics and pragmatics (also referred to as cognition). Syntax is structure or the rules by which the patterns are arranged. Semantics goes beyond syntax to find meaning in the patterns, for example, what a word in a language denotes. Beyond semantics, pragmatics or cognition concerns using patterns for a purpose—for example, how we use words or gestures to achieve our goals. Syntax is the easiest to identify, understand and communicate—for example, the rules of grammar for French and English are different and the individual patterns of symbols and syntax of Chinese are different again. Historically, the engineering of human communication has been about syntax—capturing patterns, transmitting them over a distance and transforming them back into a form we can receive. Semantics and pragmatics are much more difficult to engineer and the biggest challenge of the rapidly emerging research field of artificial intelligence. A familiar example illustrates syntax, semantics and pragmatics at work. An actor performs on stage, her words and actions are caught on camera, transformed into bits of electronic data which is transmitted by cables and satellites around the world to a receiver which sends pulses down a wire to a TV set or computer which then streams the patterns onto a screen to make a picture and sound. As the pictures and sounds change, the viewers follow the story. The signals sent down the wire and across space are what engineers call ‘raw’ data. Data captures the syntax of the patterns of sound and light of the actors and stage set and reveals how they change during the performance. Data is not information—in itself it has no semantics and no meaning. Data has to be interpreted to give it meaning and for it to become information. On stage the actor endows the words with meaning by the way she speaks her lines as written by the author. That meaning is complex and multi-layered as it is embedded in higher levels of meaning of the story and interactions with other actors as the plot unfolds. So, semantics is subtly present at the input of the syntactical patterns inserted into the TV transmission. That same semantics is available to the TV watchers at the other end as long as they have a sufficiently similar semantics or way of interpreting the syntax. But of course, since all human beings are different then our interpretations will be somewhat similar but also different—for example, some will enjoy an opera whilst others may not, some will see a particular meaning in a play whilst other make

Patterns

107

a different interpretation. If the semantics don’t engage sufficiently, then the signals become meaningless patterns.

The Syntax of Jacquard’s Loom Jacquard could not have known that his ingenious idea to communicate the syntax of a pattern of a cloth from a designer to a loom would have influence many years later and in matters way beyond weaving. His idea was to use cards punched with holes. A loom holds one set of warp threads in tension. The other weft threads are interlaced between the warp threads to make a pattern. Weavers raise some of the warp threads to form a ‘shed’—the vertical space between the raised and the unraised warps. The weft threads carried by a shuttle are then fed through the shed and then pressed into place alongside the previous threads using a batten. In a loom, the warp threads to be raised are chosen and picked up by a heddle frame. The weaver controls the heddle frame, raising and lowering as she inserts the weft shuttle. The warp threads pass through eyeholes in a series of wires (also called heddles) that hang from the heddle frame. For plain cloth every other warp is raised, the weft passed through and then the raised warps are lowered, and the lower warps raised, and the shuttle passed back. By raising not just alternate warps but different patterns of warps and by using different colours in the weft, the texture and colour can be varied to make any desired pattern of cloth. Clearly to set this up manually takes a long time. Jacquard’s idea to automate this process is shown in a simplified diagram in Fig. 6.1. A modern loom is pictured in Fig. 6.2. He had the idea of capturing the designer’s pattern in a pasteboard card with punched holes—you can see it at the top right of the diagram as a single card with a single row of holes. To keep it simple the diagram includes only one card and four warp threads with their corresponding hooks and pins—numbered 1–4. An arrangement of springs not shown in the diagram makes the pins press on the pasteboard. When there is a hole in the card the pin pushes through, as for pins 1 and 3 in the diagram. As a consequence, the hooks move forward and into the vertical path of the heddle frame. So, when the heddle frame is raised by the weaver it catches the hooks 1 and 3 but not 2 and 4. In that way, the warp threads corresponding to the holes in the card are the ones that are raised. You can see therefore that one row of holes (and no holes) corresponds to one row of the pattern of the cloth. For real applications, multiple rows of holes were punched in the cards which were then strung together in order. Jacquards loom was immediately successful and was even recognised by Emperor Napoleon. A patent was granted, and Jacquard received a lifelong pension and royalties. Around 1843, about 40 years after Jacquard, the Dobby loom appeared. It had a chain of bars with pegs inserted in holes that selected the heddle shafts to be lifted. Much later computer-assisted dobby looms have solenoids or other electrical devices to select the shafts.

108

6 Communicating

Fig. 6.1 Diagram of Jacquard’s loom attachment

Fig. 6.2 Loom. By permission of Juliet Bailey of Dash+Miller, The Bristol Weaving Mill, Bristol, UK

Signalling

109

Signalling Jacquard’s loom is a form of messaging—a designer communicating a pattern to a loom. Signals are patterns that form messages. The Crimean War between 1853 and 1856 was a turning point. It was the first war where people at home knew what was happening because a new underwater telegraph cable had been laid. It meant that news could reach London in a few hours. The editor of the Times sent William Howard Russell to report on events—he was perhaps the first-ever war correspondent. Russell described the awful conditions in vivid detail. He appealed for nursing help and so Florence Nightingale set sail for the Crimea with thirty-eight nurses. The Crimean war was also the first war to be photographed. A new era in the engineering of communications was dawning. The big change came through Michael Faraday who found ways to manipulate electromagnetic energy.

Michael Faraday Michael Faraday was the ‘father of electrical engineering’ (Fig. 6.3). He was a pupil of Sir Humphrey Davy who was a Cornish chemist and perhaps best known as the inventor of the miner’s safety lamp. Faraday succeeded Davy as head of the Royal Institution in London. In 1821 Faraday was asked to write an historical account of electromagnetism for a scientific journal called the Annals of Philosophy. Electromagnetism is all about the physical interaction between electrically charged particles and is one of what physicists call the four fundamental interactions in nature. (The others are gravity and the strong and weak interactions between sub-atomic particles.) In true scientific spirit, Faraday decided to repeat important experiments others had done. He was so stimulated by his results that he went on to make two new ingenious devices. He showed that electricity could generate physical work—in other words, he had found another form of energy. He built a device in which a wire carrying an electric current rotated around a fixed magnet and one in which a free magnet rotated around a fixed conducting wire. He made more remarkable discoveries in 1831. The first was induction, which Joseph Henry also found independently in USA at about the same time. Faraday showed that a steady current in a conducting coil had no effect on another similar coil, but a changing current created a changing magnetic field or zone of influence (we see how the idea of a field developed in a moment) and induced a current in the other loop. Induction happens when an electrical charge produces a magnetic or electrical effect in another body without any direct contact. Within a month he built the first electric generator—the ‘Faraday disc’ of copper that turned between the poles of a powerful magnet. He had demonstrated that mechanical work could be converted into electricity.

110

6 Communicating

Fig. 6.3 Statue of Michael Faraday Outside the headquarters of the Institution of Engineering and Technology in Savoy Place, London. Image by Adambro, CC BY-SA-3.0 via Wikipedia Commons

After hearing of Faraday’s work, the French instrument maker Hippolyte Pixii made a ‘magneto-electric machine’ in 1832. He produced a somewhat discontinuous alternating current (AC), i.e. one that varies in direction (Fig. 6.4). He put a coil of wire over each pole of a horseshoe permanent magnet and an iron core within each coil. He then rotated the magnet under the coils using a hand crank and thus generated an electrical current in the coils. However, as the magnet rotated so the direction of its movement with respect to the coil changed every 180 degrees. Consequently, the direction of the current also changed and so he got an alternating output. At that time direct current (DC) was much preferred so he used a see-saw arrangement, devised by André-Marie Ampere, as a changeover switch to make the current always flow in one direction, i.e. to make direct current. He had developed a very early version of a commutator—a device for reversing the direction of a current during one half of the cycle.

Michael Faraday

111

Fig. 6.4 Direct and Alternating current

The next important development was to replace permanent magnets with electromagnets to make a DC generator called a dynamo. Electromagnets are not permanently magnetic. They are iron bars with wire coils, called field windings, wound around them. When an electrical current is passed through the coils, the bar becomes magnetic. The clever part was that the dynamo was self-excited, i.e. some of the current being generated was also used to create the electromagnets. It worked because the iron bars retained some residual magnetism—just enough to enable the dynamo to get started with enough output voltage to get some current to flow in the field windings and so fully magnetise the bars. Werner von Siemens produced the first big improvement in the efficiency of dynamos when in 1856 he developed an iron cylinder armature (a rotating coil) with slots containing the field windings. After that machines were improved and made in several countries and used mainly for lighting. Siemens in Germany and England, R E B Crompton in England, Thomas Edison in USA and many others contributed applications in factories, agriculture and locomotives—the first electrical train was opened in the ‘deep’ London Underground in 1890. A new era in the generation and use of electricity was ushered in—electrical engineering moved from infancy to adolescence and a new type of engineer emerged—the electrical engineer. World War I created huge demands for electricity, but the industry had become fragmented. By 1918 in London alone, there were 70 authorities, 50 different types of systems, 10 different frequencies and 24 different voltages. In the UK an act of Parliament in 1926 created the Central Electricity Board which set up a national AC grid, running at 132 kV and 50 Hz. By 1933 there were a series of interconnected regional grids which were operating as a national system by 1938. In 1949, the grid was upgraded with 275 kV links and again in 1965 with 400 kV links. The UK grid was nationalised in 1947 and denationalised in 1989. In the USA, the Continental power transmission grid consists of about 300,000 km of lines operated by approximately 500 companies. The people running the network are professional engineers. For example, Welshman David Wright started his career designing electrical generation and distribution systems for warships. Now he is Director of Electricity Transmission Asset Management for the UK National Grid and responsible for delivering safe, reliable electricity

112

6 Communicating

supplies efficiently across England and Wales. He oversees a £1bn per year investment programme. Faye Banks is another electrical engineer who rose to become head of North East operations for the National Grid. She didn’t have an easy start in life. She grew up in the care of a local authority and a number of foster families. When she was 16, she wasn’t interested in education and left school with no qualifications. She lived in a council flat and was on job seekers’ allowance. She had to take an unskilled job but didn’t like it and was unhappy. Then one day an important machine in the factory broke down. She watched the engineer fix it and had a ‘light bulb’ moment. She knew that was what she wanted to do. But with no qualifications she had to go to college to do the exams she should have done at school. She took an apprenticeship and some extra jobs to earn enough to live on and pay college fees. She got a job with the National Grid and did so well that in 2013 she promoted to lead a team of over 100 engineers to maintain the electrical transmission network. She says that engineering has been a life-changing experience for her. National grids in all countries transmit electrical power derived from mechanical generators that are mainly heat engines—usually steam turning a turbine as described in Chap. 5. One controversial source of heat for those turbines is nuclear reaction. New and innovative ideas for creating renewable energy depend on the way we understand how energy processes work. At the time when Faraday started his research, electricity and magnetism were conceived as fluids. But his genius and ingenuity combined with extensive experimentation led him to an intuitive notion of a field. A scientific field is a region of space under the influence of a physical agency such as electricity, magnetism and gravitation. A charge or a mass in a field has force acting on it so a charge in an electromagnetic field experiences both electric and magnetic forces. He conceived the idea after noting the pattern assumed by iron filings near a magnet. His intuition was that all electromagnetic forces were distributed in well-defined geometrical patterns.

James Clerk Maxwell James Clerk Maxwell developed these intuitions into field theory. Maxwell was one of the great theoretical physicists of all time (Fig. 6.5). Maxwell derived some mathematical equations which generalised much of what Faraday had discovered about electricity and magnetism. Both Faraday and Maxwell thought of lines of force, as perturbations or disturbances of space around charges, magnets and currents. He suggested that even visible light was as an electromagnetic wave. Maxwell’s equations were a turning point in the history of science on which many others, including Albert Einstein, were later to build theories of relativity and quantum mechanics. It was, as Thomas Kuhn was later to point out, a scientific revolution. However, at the time few saw it this way. Even the famous Sir William Thomson who later became Lord Kelvin and who was Maxwell’s mentor dismissed Maxwell’s theory as ‘curious and ingenious, but not wholly tenable’. In 1904 just before he died Kelvin maintained that ‘the so-called electromagnetic theory of light has not helped

James Clerk Maxwell

113

Fig. 6.5 James Clerk Maxwell. Engraving by G.J. Stodart before 1859. Public domain via Wikipedia Commons

much hitherto’. Nevertheless, Thomson’s fame had been well earned—he was a great physicist and engineer. From 1881 he took great interest in the accumulator or storage battery invented by Camille Alphonse Faure. He collaborated with Ferranti in the design of a special winding for an AC dynamo and was a consultant to Ferranti and Crompton. He led an initiative by the British Association for the Advancement of Science to establish a common set of units on which the development of engineering science could be based. Nevertheless, it was Maxwell’s approach that presaged modern scientific thinking. By this view, the details of the basic structure of our world are beyond common sense and we must rely on mathematics as an abstract source of understanding. This has led to some mathematicians to argue that the world is mathematical. But that idea puts things the wrong way round. Mathematics is a language that expresses aspects of our understanding of how we reason rationally, and we use it to express models of our reality. Yes, we can and do describe a great deal of the natural world (but not all) using mathematics. But it does not follow logically that the world is mathematical because much of importance is not expressible that way. Mathematics is self-contained. That means the truth of any mathematical statement derives from the validity of any deductions or proofs made from its axioms. Axioms are statements that are assumed to be true. They stand or fall on their truth-likeness in reality. Consequently, the success of any way of understanding and modelling or representing the world (such as by mathematics) depends on how the model responds to tests in reality. If a theoretical model predicts something new, such as a different kind of behaviour, then we must search for it and if we find it then a test is passed. Nevertheless, no matter

114

6 Communicating

how many tests are passed there may still be a test in the future that our model will eventually fail—we can never be sure. Maxwell’s theory became powerful because it immediately passed some major tests and we found radiation at other wavelengths ranging from the longest—radio waves—to the shortest which are gamma rays. Radio waves have frequencies of around 105 Hz or lower and wavelengths expressed in kilometres and greater. Gamma rays have frequencies of 1020 Hz and wavelengths measured in nanometres (10−9 m). In between there are microwaves, infrared waves, the visible region or light, ultraviolet light and X-rays. It was completely natural to try to use electricity to send messages. The first attempts at telegraphy were in the early 1800s. 35 wires were used, one for each letter of the alphabet and numerals to send a message over a few kilometres. Each wire was dipped in a separate tube of acid at the receiving end. When current from a battery was switched into a particular wire at the sending end, then a stream of bubbles was released at the receiving end—revealing which letter had been sent. The first regular telegraph was built in 1833 when the signal was detected using a galvanometer to measure current. Samuel Morse took out a patent for his famous code based on dots and dashes in USA in 1837. The first successful commercial application was in 1839 when signals were sent from London Paddington railway station to West Drayton 13 miles away. In the UK in 1858 William Thompson patented a mirror galvanometer that was sensitive enough to detect variations of the current in a long cable. It could even detect a defect in the core of a cable. It was, at that time, the only practicable method of receiving signals over long distances. Several unsuccessful attempts were made to lay a cable across the Atlantic between 1857 and 1865. Thompson was knighted for his work on helping to lay new cables and reconnecting a lost cable in 1866. Those two cables operated until 1872 and 1877. Thomas Edison in the USA devised a way of sending two and then four messages down a single cable in 1874. By 1902 a cable was laid across the Pacific Ocean and the world was encircled. But at the same time people were beginning to explore wireless telegraphy. Guglielmo Marconi transmitted one of the first wireless signals over 6 km in 1896—still Morse code but the beginning of radio. He failed to interest the Italian government and so came to England where he eventually succeeded in sending the first radiotelegraphy (telegraph without wires) transmission across the Atlantic in 1901. In 1909 he and Karl Ferdinand Braun shared the Nobel Prize for physics for ‘contributions to wireless telegraphy’. Radiation was about to revolutionise the way we humans send messages to each other. Marconi experimented using a long pole to pick up radiating electromagnetic waves. The long pole became known as an antenna (Italian for pole) or more commonly, in the UK, an aerial. The antenna converts electromagnetic waves into electrical currents when receiving a signal and vice versa when sending or broadcasting.

Transmitting the Human Voice

115

Transmitting the Human Voice As Marconi was developing the radio, Alexander Graham Bell wanted to improve the telegraph. In doing so he invented the telephone. The telegraph was a very limiting way of communicating. Bell wanted to find a better way of transmitting multiple messages over the same wire at the same time. His knowledge of music helped him to conceive the idea of a harmonic telegraph where several notes of different pitches could be sent at the same time. By 1874, Bell had advanced his ideas sufficiently to persuade his future father-in-law to fund him. The following year he had demonstrated that different tones would vary the strength of an electric current in a wire. He just needed to build a transmitter capable of producing varying electrical currents in response to sound waves and a receiver that would reproduce these variations in audible frequencies at the other end of the wire. In 1876, Bell spoke down a wire to his assistant, Thomas A. Watson, in the next room, ‘Mr. Watson—come here—I want to see you’. A telephone is one of the simplest ways of sending data as syntax but through intonation and language it was the beginning of engineering semantics. A telephone has a switch to connect and disconnect the phone from the network, a microphone and a speaker. In general terms, a microphone is an instrument that responds to varying sound pressure waves in the air and converts them into varying electrical signals. The speaker does the reverse. So, when you speak into a telephone the sound waves of your voice are picked up, in the simplest microphone, by a diaphragm combined with some carbon granules or dust. As the diaphragm vibrates, it compresses the dust which changes resistance, and hence varies the electrical current that flows through the carbon. In a dynamic microphone when the sound waves hit the diaphragm, either a magnet or a coil moves and creates a small current. A speaker takes the electrical signal and translates it back into physical vibrations to create sound waves. When everything is working as it should, the speaker produces nearly the same vibrations that the microphone originally received. Reginald Fessenden (Fig. 6.6) decided to try using a spark transmitter—rather like a spark plug in a car engine—but with very rapid sparks. The transmitter generates signals as radio frequency waves. Fessenden modulated (changed or modified) the waves using a carbon microphone. In other words, he combined the signal from the spark transmitter with the signal from the microphone. Unfortunately, the sound it made was hardly intelligible. The problem was that the signals from the spark gap transmitters decayed over a broad range of frequencies. What was needed was a welldefined continuous oscillating wave rather than the spark transmitter wave to carry the signal from the microphone. It eventually came via the vacuum tube, perhaps better known as a valve. A vacuum tube is rather like an incandescent light bulb—but one that can switch, amplify or otherwise modify an electrical signal (Fig. 6.7). As a switch it operates like a water tap or valve that can turn flow on or off—hence, the name valve. The simplest vacuum tube is a diode, first invented by Englishman J. A. Fleming in 1904, with a filament cathode (negative) and a plate anode (positive). When the anode is

116

6 Communicating

Fig. 6.6 Reginald Fessenden. Public domain via Wikipedia Commons. Author Thomas H. White D.W. on en.wikipedia

Fig. 6.7 A vacuum tube, transister and integrated circuit. Left image: Image by Tvezymer Public domain via Wikipedia Commons. Middle image: Image by Transisto CC BY-SA-4.0 via Wikipedia Commons. Right image: Image by Ioan Sameli CC BY-SA-2.0 via Wikipedia Commons

Transmitting the Human Voice

117

positive with respect to the cathode, then the electrons move easily from one to the other—this is called forward bias. When the voltage is the other way around, it is hard for the electrons to escape from the cathode and the flow is reduced almost to zero—this is called reverse bias. So effectively diodes only allow current to flow in one direction. In 1907 in the USA Lee De Forest added a third electrode to the diode to make a triode called the audion—but it didn’t work well. In 1914, Edwin Armstrong improved the triode and showed how it could be used to amplify a weak electrical signal. The vacuum triode tube was used as an amplifier in an oscillator to produce a well-defined oscillating signal. This became the wave needed to carry the microphone signal and essential for radio, TV, radar, telephones, computers and industrial control processes. These carrier waves are modulated (changed) by an input signal—the data we wish to transmit—such as someone speaking into a microphone. There are two main ways this modulation was done. One was AM or Amplitude Modulation and the other FM or Frequency Modulation. The modulated signals—AM or FM—are detected by a radio receiver tuned to that particular carrier frequency. The carrier signal is then demodulated (decoded to find the input signal) so that the original sound signal is isolated. It is then amplified and sent to a speaker. By 1922, early radio broadcasts by the BBC were AM. Although AM is simple and robust, there is quite a lot of interference from household appliances and lighting and night-time interference between stations. It also requires high battery power. Edwin Armstrong developed FM radio in the 1930s. The BBC started FM transmissions in 1955. FM waves deliver good voice quality, are less susceptible to interference and are able to carry more information. However, the higher frequencies need a line of sight and so are interrupted by large obstructions such as high hills. The lower frequencies used for AM can travel further distances as they are reflected back from the ionosphere, whereas FM passes straight through. FM allows more bandwidth—the range of frequencies within which a particular signal can be transmitted. Nowadays, many broadcasts are transmitted digitally—using computers.

Computing Modern computing began in the late 1930s, and following Jacquard and Babbage it was entirely natural to use punched cards to input programmes and data. The first commercial computer was the UNIVAC (UNIVersal Automatic Computer) in the 1940s. By the early 1950s, many large companies were using computers for billing, payroll management and cost accounting and were exploring more complicated tasks like sales forecasting, factory scheduling and inventory management. By 1960, there were a number of companies supplying large ‘mainframe’ computers. These were institutional machines run by specialists. Individual programmers had little access but could be given ‘driving lessons’. Software for large numerical calculations enabled engineers to begin to use them and new techniques of numerical analysis began to be developed. These first computers were physically large. However, in the 1960s and

118

6 Communicating

70s, transistors and then microprocessors with integrated circuits took over leading eventually to the personal computer, laptop and handheld media player and personal digital assistant. Digital computers are binary devices. That means, at any given moment, deep inside there are ‘bits’ that are in one of two states—low or high voltage. The low voltage state is near to zero. The high volt state is at a level which depends on the supply voltage being used in a particular application. These binary levels— bits—are usually represented as (Low and High) or (False and True) or (0 and 1). The big advantage is that it is easier to switch a device into one of two known states than to try to reproduce a continuous range of values. The miniaturisation of electronics started when semiconductors and the transistor were invented—work for which William Shockley, John Bardeen and Walter Brattain were awarded the Nobel Prize for Physics in 1956. Transistors replaced vacuum tubes or valves. They enable a small voltage to control the flow of a much bigger current—just as a valve controls the flow of water in a pipe. The components are very small, very pure and solid with no moving parts and consequently are much more robust. They are used individually but they are most often found in the integrated circuits that make nearly all modern electronic equipment. By the late 1950s, these electronic components were becoming very complicated and reliability needed to be improved and costs reduced. The answer came in the form of the printed, miniaturised, integrated circuit. The impact was profound. Electronic circuits were being made on a wafer made of semiconductor material such as silicon using photographic and chemical processing in highly specialised facilities. These clever devices are another example of the creative artfulness or aesthetic of function. A few hundred transistors could be formed on one silicon chip and linked together. In the 1970s, the number of transistors integrated onto a silicon chip doubled every couple of years—a phenomenon that came to be known as Moore’s Law after Gordon Moore who co-founded Intel Corporation in the USA. Integrated circuits are found now in almost all electronic equipment. In 2009 Intel unveiled a prototype Singlechip Cloud Computer (SCC) that has 1.3 billion transistors on a silicon chip the size of a postage stamp. As of 2016, the largest transistor count in a commercially available single-chip processor is over 7.2 billion—the Intel Broadwell-EP Xeon. In other types of integrated circuits (IC), such as field-programmable gate arrays (FPGAs), Intel’s Stratix 10 has the largest transistor count at over 30 billion. One of the very basic devices for manipulating information deep within an IC is a logic gate. They are found, for example, in flip-flops as memory stores, counting and other applications. Computers and computer networks pass messages to each other using a set of communication protocols or rules for governing the format of messages. Looking vertically down into the subsystems of your computer is a layer of interconnected semiconductors working together to make the logic gates in the next layer up. The gates work together in a third layer to make flip-flops and other devices making up patterns of bits and bytes. Next are the registers, memory cells and arithmetic units made of connected flip-flops and all put together into integrated circuits that make up the brains of a computer—the microprocessor or central processing unit

Computing

119

CPU. The top hardware layer contains several other connected components such as memory, disc drives, monitor, power supply, keyboard and mouse all working together to make up the characteristics of your particular machine. The next layer is the first software layer with the set of instructions that make the microprocessor work. The lowest of these are the machine languages; the next layer contains the programmable languages such as Fortran, C and the Object-Oriented Programming (OOP) languages such as Java and C++ that use message passing between software objects. Then come operating systems such as Windows and finally applications such as word processors, spreadsheets, web browsers, emails and Internet protocols and connections. In your computer, the Internet protocol has four layers: (i) the data link layer (for example, an Ethernet may handle communications between physical network components), (ii) the IP (or Internet protocol or network) layer which routes packets of information across networks, (iii) the transport layer between networks and (iv) the application layer which is the highest layer that sends and receives data for particular applications such as HTTP—hypertext transfer protocol. Every layer in your computer will be successful if the lower layers are successful. Some of these processes will be necessary (e.g. electrical supply) and some will be sufficient (e.g. one of two alternative microprocessors used to create redundancy). We can look outwards from your computer both horizontally to the servers to which you are connected directly and upwards to the Internet itself. The way the Internet works is analogous to the posting of ordinary letters and parcel mail. Data, such as a web page, a downloaded file or an email message, travels over a system known as a packet-switching network just as a letter travels through the international mailing systems. First, the data is broken up into packages or packets—analogous to a letter or parcel. Your computer Internet protocols attach a wrapper to each packet that includes information on the addresses of the sender and the receiver and the place of the packet in the message. There is no direct link between sender and receiver. Each packet is sent separately to a router on the packet-switching network. The router is analogous to a mail sorting office. It matches the information in the packet with its own information and then passes the packet onto another router. That next router then passes the packet to another router and so on. The packets are travelling at the speed of light, so they take only seconds to get around the world. Eventually, each packet arrives at the receiver but often by different routes and at different times. The protocol layers at the receiver reassemble the packets into the original data before sending it to the monitor of the receiver. The routers are computers containing algorithms (rules) that receive packets and choose where to send them next. The algorithms are constantly being updated to ensure the best routes are chosen. The packet-switching network is therefore effectively a set of processes that are interacting by passing messages to each other. The complicated behaviour of the Internet emerges from the layers of interacting simplicity. Each operation in the system is on its own quite simple, but when connected to all of the other simple processes a complicated but rather elegant whole is created—another example of an aesthetic of function. In 1971, Ray Tomlinson wrote the first email programme on the Advanced Research Projects Agency Network (ARPANET) which was the precursor to the

120

6 Communicating

Internet. Previously, mail could be sent only to others on the same computer. Tomlinson worked out a way to send mail between users on different computers connected to ARPANET. He used the @ sign to separate the username from the name of their machine and so changed the way people communicate Telecommunications went through another transformation in the 1980s through optical fibres. Glass or plastic fibres can carry more information than other forms of communications. They are better than metal wires because they offer higher bandwidth, are lighter in weight, have low transmission loss, are less sensitive to electromagnetic interference and are made from cheap material. The fibre-optical Internet is doing for computing what the AC network did for power distribution because where the equipment is located is not important to the user. The lone PC has given way to cloud computing where computing is a utility—just as is electricity. Computing is sold flexibly on demand—users can buy as much or as little as they want. Interactive websites don’t just provide information—they facilitate interaction between users as in social networking sites. The social impact of these technologies is as great as steam railway or the automobile.

Semantics and Pragmatics Our success in communicating the syntax of messages carries the risks of technological triumphalism that I referred to in Chap. 3 and an underplaying of the difficulties of semantics and pragmatics. The way we humans assign meaning to a pattern depends on our individual experience and learning—though of course much is shared. As soon as we are born (and perhaps earlier1 ), we perceive patterns through all of our senses, and we start to try to make sense of them as best we can. We learn from our parents and others to interpret sights and sounds as spoken language. We learn to read words as written language. We feel the touch of other people, hair, feathers and the fragrance of flowers and perfume. We appreciate paintings, music and birdsong. All of these sensations are processed by our active brains as patterns of electrical and chemical signals. Neurons are connected via synapses in massive networks. Sensations trigger patterns which neuroscientists can now scan. In that way, they are beginning to understand how our brains respond to stimuli. The philosopher Karl Popper explained the way we perceive, sense and share, this way. He said that we effectively inhabit 3 worlds. First world 1 is reality—the actual physical world of which we are a part—in accordance with the principle of Part. Second, he said that we can only make sense of world 1 through world 2—our own subjective world of mind. World 2 is where we think about the things which we cannot share with anyone else. For example, you can describe the pain in your stomach but no-one else can feel it as you do. You can say what love feels like but no-one else can feel it as you do. Third, he said that we also try to make sense of world 1 (and world 2) through world 3—the world of our shared experiences. That is the world of objective data. We all see the same moon; we talk through shared

Semantics and Pragmatics

121

languages and we share information and knowledge. World 3 contains everything we share—the theories of science as well as fictional literature. A useful way of thinking about world 3 is that it consists of all the books in all of the libraries in the world and information on the Internet. Some things in world 3 are true and some are false—science, said Popper, is shared information that is testable. The German philosopher Arthur Schopenhauer wrote in the early nineteenth century that we cannot know the world itself; we can only know our experience of it. Our understanding is demonstrably partial and incomplete. Certainty is available only through faith—we cannot logically and rationally prove everything in the vast cosmos, or indeed in our own world, with total and absolute certainty. But we can see the intimate connection between ourselves as individuals with our social systems and with the physical and natural world. This way of understanding requires us to keep constantly in mind that Popper’s worlds 2 and 3 are contained entirely within world 1. In other words, there is a reality, but it is only available to us and we can only try to make sense of it in our own subjective minds of world 2. But then we share and test our own individual understanding with others in the objective world 3. We form theories in worlds 2 and 3 but we act in world 1—for example, to carry out an experiment. Engineering communication is about finding ingenious ways of creating patterns in world 1 which make sense to us in worlds 2 and 3. The terms subjective and objective are often used confusingly. They are important in engineering, and especially in engineering communication because the difference between data and information is critical. For example, the Google search engine looks for patterns in data not information (though that is changing as we shall see in a moment). That is why it sometimes gives inappropriate answers. If you search the word ‘ham’ you get items about food as well as the UK football team West Ham United. The search engine is looking to match patterns in web data that match patterns in your search input data. Researchers are working on ways in which computers can look at patterns of syntax and hence begin to provide answers that make more sense but despite considerable advances that is still work in progress—artificial intelligence (AI). In the meantime, Popper’s interpretation helps to make the differences clear. The word subjective means things that belong to the world 2 thinking mind of an individual. The word objective refers to things outside of the mind of any one individual and hence available to everyone—the content of world 3. The crucial point to note is that objective knowledge may be true or false, fact or fiction. Science is world 3 testable information. That does not imply that information that is not testable is not valuable—fictional stories, myths and religion are not directly and precisely testable, but they can reveal important understanding of the human condition as the following BT story reveals.

122

6 Communicating

Perspectives In the 1990s, a group of BT senior executives was meeting in an office off Fleet Street London.2 Reportedly, the meeting changed the company. It was an example of pragmatics in action. The Director of Communications at BT told the group they were not in the telephony business, nor were they a technology business or the business of installing equipment. In his view, BT was in the business of ‘Reciprocated Confidences’. Not surprisingly the executives were confused. They needed him to explain. As they talked his ideas over, they began to see their business from a new perspective—when people exchange confidences, they communicate better and, in turn, form deeper relationships. But how were they to make that idea meaningful to their customers? They brought in a communication agency. The agency suggested that BT needed to change the attitude of the (mainly male) bill paying public. Women are more likely to share confidences than men—blokes don’t do that and some don’t approve. The agency’s final report was a hefty tome but was topped with four words. It’s good to talk. The advertising campaign that followed delivered an incremental £5 billion over 5 years—an astonishing result. Good communicators know that active listening is also important—but that might have been perhaps not such a catchy slogan. Active listeners show what they do. They paraphrase to demonstrate concern and understanding, to create rapport and give nonverbal cues such as nodding, eye contact and comments like ‘I see’, and ‘I know’. In this respect, the spoken and written word are quite different. Once spoken words disappear only half-remembered. It is active information—interactive, usually short range and ethereal. Written words are not the same—they can be found again. Some call this passive information since we don’t interact with it—we receive and do not change it.

Passive and Active Information We tend to think of all information as passive—rather static like the patterns on our woven cloth, in the written word and in our numbers. Passive information is material—for example, reading a book. The first writing was passive, cave paintings, static clay and wooden block type for printing that developed into moveable metal stereotype and now held in computer memory. Passive information is constant and unchanging—we don’t expect the words in our favourite novel to change each time we read it. Counting is also passive information—simple and stable. We count using the decimal system, but computers use a binary system. Active information is quite different. It is in our brains, the spoken word, and in our increasingly dynamic world of changing social media. Active information is material (for example as patterns in our brains) but always changing because of almost continuous interactions between the sources. Coping with the pace of

Passive and Active Information

123

change is increasingly challenging since it may be ethereal, misleading as it strives to command our attention, easily misinterpreted and worst of all malicious.

Big Data Becomes Active ‘Big data’ (extremely large data sets) and even smaller miniaturisation of components are set to pose even greater challenges of unintended consequences—again with risks of technological triumphalism if we are not careful to separate the hype from the reality. Big data is moving from passive to active. But how big is big data? Michael Gillings in Sydney and colleagues have compared the amount of data and information available through the Internet with the natural world. Estimates of the storage capacity of the Internet are around 1024 bytes3, 4 and growing fast compared with estimates of information in living things (DNA) of around 1031 bytes.5 Some researchers forecast that the gap could be closed within 100 years at the present rate of progress. That is speculation but certainly the increase in the rate of progress is impressive. If we start from the time of the first living matter, around 4.25 billion years ago, there were only RNA genomes (genetic material) and then DNA eukaryotes (cells with nuclei) appeared after about another billion years. Multicellular organisms with a nervous system took approximately another 2 billion years to develop. Neural systems capable of language came around 100,000 years ago and writing around 3500 BCE. It took approximately 4,500 more years to get to the printing press and in 1980 only about 1% of information was in digital format. Today most information is digital. The vast changes ahead and the numerous unknown unknowns and unintended consequences are part of the long route from primitive signalling to social media.

Artificial Intelligence The real challenge of creating computers that can interpret and use data as information and knowledge (semantics) and then act on it (pragmatics) is being tackled by Artificial Intelligence (AI). There isn’t a commonly agreed definition but in essence AI is performing tasks that normally require human intelligence, such as visual perception, speech recognition and decision making. But intelligence is a difficult idea—for example, a thermostat can perceive and adjust the temperature in your house, but we would not normally label it as intelligent even if you can control it remotely on your mobile cell phone. The thermostat is quite different from AI controlling a drone to select and engage a target in a war zone without any human intervention. New media create a lot of hype around AI—they report companies replacing workers with AI and algorithms are doing better than doctors at diagnosing complex conditions. We are told that machines are ‘powered by AI’ can do everything from forecasting the weather to ordering a pizza. As a result, there are some misconceptions about what AI can and cannot do.

124

6 Communicating

John Launchberry of the USA Defence Advanced Research Projects Agency (DARPA) is more level-headed. He describes three waves of AI which he calls handcrafted knowledge, statistical learning and contextual adaptation.6 Currently, he says, we are enjoying the fruits of the first and second waves and embarking on the third. The first wave of handcrafted knowledge was about using expert knowledge about a specific and quite precise domain to synthesise some logical rules that could be programmed in a computer. The computer could then be used to imply results from those rules. Typical applications are playing games like chess, scheduling product deliveries and even rules that govern how much tax one should pay. These systems are good only in well-defined and well-understood specific situations. They are able to perceive measured attributes such as the dimensions of an object, and reason using logical rules in algorithms, but unable to handle uncertainty, or to learn and abstract models at higher levels of understanding. A good simple example already mentioned is a house thermostat that works through deterministic (cause and effect) rules— if A then B else C. For example, if (temperature is above target) then (turn down heating) else (do nothing). Since it is deterministic, we always get the same output for the same input (unless something fails). Systems like this readily fail in more complex situations that may stray beyond the rules programmed in—for example, these kinds of systems failed in early tests to control driverless cars because they could not distinguish hard objects from shadows. The second wave of AI statistical learning has led to the recent very impressive and fast-growing advances in ‘big data’ and computer power based on so-called Artificial Neural Networks (ANNs). They are being used to read handwriting, interpret vocal sounds and process natural language. Indeed, you may well have the software on your computer or mobile cell phone that you can use to ask questions, or a smart speaker in your home to play music or control your central heating. Systems like this are the basis of the software that will control the first driverless cars. Other applications may be forecasting time series such as stock market prices, integrating lots of sensor data such as humidity, barometric pressure, dew point and air density into one measure of air quality, and detecting anomalies such as faults in a product. ANNs developed from an idea that Frank Rosenblatt had in USA in 1957 for a machine that can ‘learn by doing’. An ANN consists of an array of weighted connected nodes and is, in effect, a mathematical function. It creates a relationship between a set of inputs and a set of outputs. But it isn’t a traditional mathematical formula, like y = 3x + 7, or more formally y = f(x), which maps values of inputs x to outputs y, rather it is a pattern of nodes and links as shown in Fig. 6.8a. The idea arose through an analogy with the biological connections between neurons in our brains—so the nodes are often called artificial neurons (AN). Of course, ANs and the connections between them are much simpler than the neural networks in our brains. ANs are of three types, input, output and hidden (Fig. 6.8a) and each has a value. The overall connection between inputs and outputs is called a mapping. A mapping is made up of a multifaceted pattern of intermediate connections between the input and the hidden internal ANN nodes which are, in turn, connected to the output ANN nodes. The nodes contain rules through which the values of each node are calculated by combining the weighted values of the nodes to which it is connected.

Artificial Intelligence

125

Fig. 6.8 a An artificial neural network. b An artificial neural network with no hidden layers

As a consequence, the values and weightings may change as the ANN ‘learns’. That learning is a ‘training’ process in which examples of known inputs with known outputs are fed in. An ANN with lots of layers of hidden ANs is called a ‘deep’ learning ANN—it is a type of emergent layering as per the principle of Part. We can get a glimpse of the structure of an ANN with a simple example from a logic gate in a computer—although computers don’t usually have an ANN for this purpose and the process isn’t one that learns. Imagine an ANN with three ANs only as in Fig. 6.8b. We want to represent the logical disjunction ‘OR’ in X = A OR B. For example, X = (Joe is alive), A = (Joe is breathing) and B = (Joe has a pulse). A and B are input ANs and X is an output AN, and there are no hidden ANs. A is either on (1) or off (0) as is B. The meaning of X = A OR B is that X is on (1) in all cases except when both A and B are off at the same time. We can make an ANN to

126

6 Communicating

do what we want by allocating random non-zero positive weights of w1 to the link from A to X and w2 to the link from B to X. Then we use a rule that X is on (1) if A* w1 + B* w2 > 0 otherwise X = 0. In that case, X is always on except when A is 0 and B is 0. In other words, the ANN says that Joe is alive when he is breathing or has a pulse but if he has neither then he is not alive. A key reason for using an ANN, unlike the ‘OR’ example, is that it can learn. But to do that most applications require very large ANNs with thousands of ANs and maybe 20 or more hidden layers. They are sometimes called adaptive because the internal structure changes according to the data flowing through it. The term ‘deep learning’ used to describe the process perhaps owes much to ‘techie’ triumphalism rather than the actuality. Indeed John Launchberry6 has called ANNs ‘spreadsheets on steroids’ because so much data is needed. So-called supervised learning occurs when the ANN is fed with examples of inputs and outputs, such as various photographs of faces and their names. Each photograph is saved as pixels, say 224 × 224, and each pixel is an input AN. In practice, various techniques maybe used to reduce the number of ANNs by extracting a limited set of features (such as the distance between eyes) to use as ANs or to reduce the number of pixels by a selection process that takes representative samples from a small region of the photograph. Each input AN is given an intensity value depending on the lightness/darkness of the pixel it represents. The output AN is given a value which is the name of the person whose face is in the photograph. Using the first photograph the network ascribes some random weights to the links between ANs. Then a second photograph, with a known output name, is scanned and its intensity values for the ANs are combined with the weightings from the first photograph to make a prediction. If the prediction is wrong, then the guessed answer is combined with the known answer using a set of pre-programmed rules and the weightings and nodal values are adjusted. A third and many more (perhaps as many as 50–100,000) subsequent photographs are used so that through continuous adjustments the ANN is ‘trained’ and adapts to the new information. Eventually, the ANN is able to predict, with a high degree of accuracy (>90%), the name of a person it has learned about from previous photographs from the scan of a photograph of that person it has never seen before. The learning works because the training adjustments are adjusted or tuned so that the data space (the graph of the many attributes or features) is stretched and squeezed so that individual clumps or clusters of data can be cleanly separated—for example, the nose of a face. This second wave of AI is good on perceiving and learning but very poor at abstracting and reasoning. Also, the ANNs can occasionally come up with bizarre answers that no human would countenance—for example, identifying a toothbrush as a baseball bat. The systems can also be tricked into giving wrong answers by overlaying invisible distorting patterns designed by a malevolent expert. Another issue is that answers are entirely dependent on the input training examples so any bias or distortion, intended or unintended is built in—for example, gender or racial or other cultural assumptions. Although the success achieved is impressive, the need for large data sets is not ideal especially when you think that human beings can recognise a face after one glimpse—the ANN is far from mimicking human intelligence. As

Artificial Intelligence

127

one acerbic commentator said if a computer ANN had to learn how not to get killed by a car, then it will have to get killed before knowing how not to get killed. Unsupervised learning is different from supervised learning because the ANN looks for clusters of ANs based on rules and measures of proximity. A third type of learning is call reinforcement learning and used in robotics when a robot performs a task and observes the results. John Launchberry’s third wave of AI as contextual adaptation is in its infancy. The idea is to build explanatory models that can identify a face by saying it has certain features (much as we might), for example, dark hair, bushy eyebrows, blue eyes, rosy cheeks, etc.—in other words, the system should be able to say why it thinks the face is of a particular person and what is more do it with only a few training examples. Such a system would begin to be good at perceiving, learning, reasoning and perhaps even abstracting—but that will require much more research. In 2017, Geoffrey Hinton (who is often called the ‘father’ of ANNs) has suggested that the ANNs with a single number be grouped into what he calls a ‘capsule’.7 This will contain a vector of values and do some internal computation and output a compact result to other capsules. The idea could lead to being better able to identify relationships between features—for example, how the eyes, nose and mouth of a face appear in relation to each other. He also suggests that the new method will enable the ANN to recognise a face from a different angle than seen previously with fewer training examples. If we are ever to feel we can trust AI systems to replace humans as professionals, then they will have to have the capacity to explain themselves—at the time of writing no systems exist that can do that, but research is ongoing. When that happens, we might ascribe some form of understanding to them at a level commensurate with their perceived capability. Currently, if a doctor diagnoses a patient with cancer then the patient expects the doctor to be able to communicate why he thinks that and state what evidence he has to form his judgement. Similarly, we must expect AI systems to give us reasons for any decisions to justify at a level commensurate with that required by the human asking the questions but also and necessarily when human safety is at stake, at a level commensurate with a duty of care under the law of tort. Otherwise, these systems are simply opaque ‘black boxes’ which we can use as evidence to support the human decision maker but no further. As currently formulated, there is no way can we trust AI systems to make any important decisions about public safety, such as the design of a bridge or diagnosis of cancer, other than as advisors to human decision makers. But there are some important and very specific exceptions such as driverless cars. Vehicles have the potential to injure and kill people, but we don’t expect a human driver to justify why he changed gear or turned left at a particular junction. The acid test for vehicles is safe performance. The public will expect driverless cars to reach their destinations safely (as well as comfortably and efficiently, etc.). But they will have to show a much higher level of safety than cars driven by humans. The reason is that the people being transported will feel that any control of what is happening has been taken away from them. It is well known that under those circumstances people expect higher levels of safety.

128

6 Communicating

Women in AI Professor Dame Wendy Hall took up the challenge of engineering the Internet and tackling its semantics. She graduated in mathematics and became her university’s first female professor of engineering in 1994. She has shattered some glass ceilings by her work on a number of national and international organisations to promote the role of women in engineering, science and technology. Wendy is now the Director of the Web Science Institute at the University of Southampton UK. Web science is the study of large-scale social–technical systems like the world wide web. Early in her career, she realised that the web is as much about people as it is about the computer network. Wendy was one of the first computer scientists to undertake serious research in multimedia (i.e. text, audio, images, video, hyperlinks, etc.) including digital libraries (electronic media), the development of the semantic web and web science. The semantic web (also called the linked data web) attempts to create meaning by exposing, sharing and connecting data using models of extra data called metadata. She and James Hendler wrote8 : ‘As we see the increasing emphasis by politicians and funding bodies on the economic advantages of more students entering the STEM area, there has been a concomitant tendency to treat the social sciences and humanities as if they are somehow of lesser value. However, not only is the Web a network of machines, it’s a network of billions of people throughout the world interacting together in never-before-possible ways’. Judy Wanjiku Maina is less well known but her work as a telecommunications engineer in Nairobi, Kenya is important. She oversees various government projects in digital literacy—for example, making sure laptops are delivered to public primary schools and properly maintained. She says engineering is fast and vast with endless opportunities to learn and grow. She loves to see how different components connect and then work together to become a new machine. She is active in a group called WomEng (Women in Engineering)9 which she says has inspired her, helped her to learn about herself and given her extra skills. WomEng is promoting engineering for females largely in Africa. Its aims are to create greater awareness of engineering, to attract and retain women in engineering, to provide mentors and to develop skills including leadership, and innovative problem solving. Naadiya Moosajee was one of the founders of WomEng. She says she still get gasps or looks of surprise when she tells people what she does. She explains the small proportion of women engineers as ‘the pipeline challenge’. Women and girls are lost at all stages along the engineering talent pipeline. At the beginning, girls are not aware of engineering careers. There is still a perception that it’s something for the boys. People are surprised that she is a civil engineer because she is small. They seem to think that a lot of heavy lifting is involved. She frames it as heavy lifting with the mind—especially through the use of big data, AI and robotics in construction. She says that most of the female engineers she knows have entered the profession in one of three ways: their dad, uncle or another man was an engineer; a teacher said they were good at maths and science; or ‘sheer luck’. She became an engineer via sheer luck and that is why she started to campaign for STEM (science, technology,

Women in AI

129

engineering and mathematics) subjects in schools. She calls WomEng the ‘pinkhard-hat transformation’. The founders of WomEng took the stereotypical hard hat seen on many construction sites and made it bright pink—fighting a stereotype with a stereotype.

Getting Smaller Big data and AI have been driven by dramatic miniaturisation of computer components since the invention of the transistor and integrated circuit—and it is ongoing. DNA computing is a research topic based on molecular biology rather than silicon chips and uses the four key components of the DNA molecule. Quantum computing replaces the bits of 0 or 1 in electronic computing by quantum phenomena at the atomic level. This technology may provide as big a change in scale as that from valves to transistors though getting from theory to practice is not easy. Nanotechnology works with very small particles that are comparable to the size of a biological cell. A nanometre is one billionth of a metre or one millionth of a millimetre. A cell measure 10–100 nanometres and a gene may be 2 nm wide with a length similar to a cell. Magnetic nanoparticles offer some attractive possibilities in biomedicine. For example, they can be made to bind to a cell giving that cell a tag or ‘address’—in a somewhat similar manner to a computer memory address. They can then be used for imaging, tracking or as carriers. As they are magnetic, they can be manipulated by an external magnetic field. They can be made to deliver a package, such as an anticancer drug to a tumour. The particles can be made to resonate to heat up, and hence kill tumours or act as targeted agents of chemotherapy and radiotherapy. The research could lead to better tools for screening different diseases in a non-invasive and accurate way, and for administering therapeutic agents safely and effective with fewer side effects. Another technique called photodynamic therapy uses light to destroy cancer cells. The patient is injected with a photosensitizer which accumulates in a tumour. Using a precise laser, the tumour is then targeted with light of a wavelength that will be absorbed producing a drug that kills the cell. The big breakthrough will be to translate innovations in the laboratory into commercially viable medical products. Energy is a property of material objects in world 1 seen through ‘the eyes’ of world 3. Data is abstract and immaterial in world 3. Data has to be embodied in material form to be transmitted by a process in world 1 using energy. At root the embodiment of digital electronic information consists of patterns of 0 s and 1 s and at face value passive. However, because there is now so much data transmitted at the speed of light, and because AI is growing ever more powerful, digital data is now most often active. The impact of some forms of modern social media is to move passive information (as in newspapers and books) to active information through Twitter and Facebook. Data and information are becoming much more complex and the risks of unintended consequences abound. We are moving into an age of systems as we begin to understand how complexity can emerge from simplicity. Later (Chap. 9), we will return to the grand challenges of big data, the human issues surrounding

130

6 Communicating

social networking, cybersecurity and the problems for our business, political leaders and opinion formers as they struggle to deal with what is happening. Before we do that, however, there is one aspect of human behaviour that we have only touched upon—our ability to misunderstand each other and to fight. How has engineering both helped and hindered our ability to make war and peace? What are the risks of using AI in a war by remote control? End Notes 1. Babies in the womb exposed to a melody remember and respond to according to a study by Finnish researchers. See http://journals.plos.org/plosone/article/ file?id=10.1371/journal.pone.0078946&type=printable (Last accessed February 2019). 2. BT Its good to talk https://www.campaignlive.co.uk/article/its-good-talk-storybehind-campaign/938629 (Last accessed February 2019). 3. Gillings, MR., Hilbert, M., Kemp, DJ., (2016, Information in the Biosphere: Biological and Digital Worlds Trends in Ecology and Evolution, March, Vol. 31, Issue 3 See http://www.martinhilbert.net/wp-content/uploads/2016/02/ Information-in-the-Biosphere_TREE-final.pdf (Last accessed February 2019). 4. Satoshi Matsuoka, et al. (2014), Extreme Big Data (EBD): Next Generation Big Data Infrastructure Technologies Towards Yottabyte/Year Supercomputing frontiers and innovations, 1(2): p. 89–107 See http://superfri.org/superfri/article/ view/24 (Last accessed February 2019) Note that 1024 is 10 with 24 zeros—a very large number called a yottabyte. 5. Landenmark, HKE., Forgan, DH., Cockell, CS. (2015) An Estimate of the Total DNA in the Biosphere, Plos Biology, June 11, 2015 See http://journals.plos.org/ plosbiology/article?id=10.1371/journal.pbio.1002168 (Last accessed February 2019). 6. Launchberry, J., A DARPA Perspective on Artificial Intelligence, Lecture at https://www.youtube.com/watch?v=-O01G3tSYpU (Last accessed February 2019). 7. Sabour, S., Frosst, N., Hinton, GE. (2017) Dynamic Routing between capsules, 31st Conference on Neural Information Processing Systems (NIPS 2017), Long Beach, CA, USA. See https://arxiv.org/pdf/1710.09829.pdf (Last accessed February 2019) See also Pechyonkin, M (2017) Understanding Hinton’s Capsule Networks. https://medium.com/ai%C2%B3-theory-practice-business/ understanding-hintons-capsule-networks-part-i-intuition-b4b559d1159b (Last accessed February 2019). 8. Hendler, J, Hall, W. (2016) The Science of the World Wide Web, Web Science, 354, (6313), p. 703. 9. For WomEng see http://www.womeng.org/ (Last accessed February 2019).

Chapter 7

Fighting

Unprepared running scared, unrepaired well impaired

Literary Engineering Engineering is not often associated with fine literature so when I was asked to speak about Engineering at the ‘Ways With Words’ Literary Festival in Dartington Hall in Devon, England in 2013 I readily accepted. I soon found out that a colleague Professor Bill Doyle was due to speak immediately before me about his book on the French Revolution. At lunch, my wife Karen and I were enjoying talking to Bill when I heard a loud voice declare ‘Hello Professor Blockley. Do you remember me—you taught me about structural engineering’. The voice was James Crowden who I didn’t immediately recognise as he had changed somewhat from his student days. But I did remember his name and soon recalled his visage. I couldn’t remember the exact date, but he told me that he graduated from the University of Bristol as a civil engineer in 1976. In those days, I was a very young and rather naïve lecturer with a lot to learn about students, but I didn’t admit to that. As we talked over lunch James told me that he did not become a professional engineer. He joined the army and served in Cyprus. He travelled in Eastern Turkey, Iran, Afghanistan and northwest India. In 1976–77, he spent a winter on the northern side of the Himalayas, in the remote Zangskar Valley in Ladakh. He said that the experience had developed in him a lifelong interest in agriculture and Buddhism. He decided to study ethnology (a branch of anthropology) at Magdalen College and the Pitt Rivers (archaeological and anthropological) Museum, Oxford. After that and for the last 20 years, James has worked in North Dorset and South Somerset as a shepherd, sheep shearer, cider maker and forester. Now he is an author, publisher and poet. He said that his choice of manual work was deliberate because it gave him a deeper understanding of the © Springer Nature Switzerland AG 2020 D. Blockley, Creativity, Problem Solving, and Aesthetics in Engineering, https://doi.org/10.1007/978-3-030-38257-5_7

131

132

7 Fighting

landscape. He remembered Afghanistan as a wild, but very beautiful country. He said ‘I think of the Lapis Lazuli mines in the Munjan valley, crossing the Hindu Kush on foot with another Bristol civil engineer Peter Grace, via the Kotal Ramgul and then finding the valleys of Nuristan at exactly this time of year, rich with apricots and mulberries, apples, walnuts and wild grapes. But I also think of the old man I found cradling an 1878 Snider Enfield percussion cap rifle captured from the British in 1878 and still in use 100 years later. True, we were held hostage by villagers at one camp for 2 days but that was because they mistook us for Russian geologists, and it was no worse than being held up by Kurdish communist bandits on Mt Ararat. The idea that Afghanistan should be hit with cruise missiles fills me with as much horror as the destruction of the Buddhas of Bamian’. James told me that he was to chair my session at the Festival. I asked him directly if his engineering degree had helped him in his career outside engineering—especially as a poet. After all, one of the great nineteenth-century engineers Thomas Telford was also a published poet and Robert Stephenson’s grandchild was the Scottish, novelist, poet and travel writer Robert Louis Stephenson. Cuban American Richard Blanco who read his much-praised poem ‘One Today’ at the swearing-in ceremony for President Barack Obama in 2013 also graduated in Civil Engineering in Miami and practiced as an engineer since 1991. James replied unhesitatingly—yes, very much so. So, I asked if he would say that when he introduced me to the festival audience—and he did. Neither of us remember his exact words but it was something like this. ‘I studied engineering with Professor Blockley, but I haven’t pursued a career in Engineering – I was in the army and now I am a poet and publisher. I feel that the structure of engineering is important, and I certainly found it useful with publishing projects, anticipating problems, etc. as well as developing an open mind to unusual situations. I went on to study anthropology and then was involved in the study of the social effects of a road going into a Himalayan Valley. Engineers are never usually around long enough to see the effects of their structures i.e. roads, bridges, dams etc. and the effect on the community, socially, psychologically, and politically and even genetically if new groups can intermarry. And anthropologists/sociologists rarely used to study societies in flux or change, though I hope they do now. As with tall buildings, the social structure has to be strong but flexible. Allowable movement up and down the social scale as well as a bit of sideways movement. I think there are many more analogies that could be brought into play. Social structure is a very important part of anthropology and when studying at Oxford I was often brought back to thinking about the lectures on the engineering of structures. And how often the craftsmen, the proto engineers in society were either up or down the social scale. The grounding in civil engineering was very, very useful particularly with problem solving’. James is a living example of someone who has used his engineering education to prepare him for quite a different life. Engineering is not simply vocational, in the sense of just a job, rather it is about the intimacy of knowing and doing. Many people take for granted that the ‘academic’ is superior to the ‘vocational’. We have lost the idea of a vocation as a calling or strong inclination rather than simply a job. Academic is typically linked to intelligence, cleverness, getting a university education

Literary Engineering

133

and having wide applicability. Science is viewed as being academic. Vocational is typically linked to being practical, working with your hands, being an apprentice if you are lucky and having only a limited scope. Engineering is viewed as being vocational. In an important sense, this is a self-fulfilling assumption. We forget at our peril that neither would exist without the other. Being interested in things ‘practical’ does not imply an inability in ‘theory’. Likewise, being interested in things ‘theoretical’ does not imply an inability in the ‘practical’. Engineering is a discipline that develops a capability for identifying and solving problems, for thinking creatively whilst being able to make decisions based on defined criteria. As an engineer, you become numerate and not frightened of mathematics or the advanced science and technologies of modern life. You appreciate the dependability of information, of scientific modelling and the need for judgement based on a professional duty of care. Since your solutions will be tested by ‘Mother Nature’ you must be honest and that tends to spill over into personal relationships—almost all engineers I know are honest and straightforward WYSIWYG (What You See Is What You Get). You identify personal and team values and develop a strong sense of moral purpose and ethical behaviour. Curiously perhaps James graduated in civil engineering and joined the army to be a military engineer. Nowadays, civil engineering is that branch of engineering that deals with buildings, bridges, water and the environment. Military engineering is about designing and building military works, maintaining lines of military transport and communications and the logistics behind military tactics. Modern military engineering is not therefore non-civil engineering. Rather, it is a particular application of all forms of professional engineering to conflict and warfare. For example, Suzanne Stamford is a platoon commander responsible for electronics in REME (UK Royal Electrical and Mechanical Engineers). She was sent to Iraq to lead 50 men responsible for controlling the digital systems of the Battalion. She had to make sure that technicians possess the relevant trade skills to maintain and operate equipment. As well as training she was responsible for maintenance including, for example, a fleet of robotic ‘wheelbarrows’ used to inspect or detonate explosive devices. Jayne Bryant is another woman engineer involved in the Defence industry. She says that she took up engineering by accident. Just near where she lived as a girl GEC Marconi ran a course for aspiring software engineers. They offered to pay her to take the course—too good to refuse—so she signed up. Now she is the Engineering Director Defence Information for BAE—leading a team of around 600 engineers. Jayne is responsible for delivering defence information, products and services in the total lifecycle from first concept to last disposal. She has to help ensure BAE’s products and services meet customer needs and all regulations. She says she used to get most satisfaction from her own achievements but now gets her biggest kick out of seeing others do well. She says the ‘hard hat’ image that is often portrayed for engineers is not helping to encourage women into engineering. She says that she has never felt at a disadvantage by being a female engineer and has managed to fit family life into her career. She feels strongly a need to engage with girls to explain what a great career engineering is. In 2019, Sue Gray at the age of 56 and a Royal Air Force engineer was promoted to the rank of Air Marshal and consequently the most senior female military officer

134

7 Fighting

in the British Armed Forces. She is to be Director General of the Defence Safety Authority overseeing regulation, accident investigation and safety. Previously, she has worked on VC10 Transport aircraft, helicopters during Gulf Wars, and has led the delivering of engineering and logistics support for fighter, training and remotely piloted aircraft. She says ‘I have been fortunate to have a job that was more of a way of life…… As an Engineer Officer…. I am part of a team, who succeed together……. through some challenging situations on operations and back at home’. The central question for this chapter is, what is the role of engineering in warfare? Whatever your ethical stance and like it or not, why have engineers created ever better ways for us to kill each other? From the ancient trebuchet (catapult) through bows and arrows, swords, gunpowder, guns, cannon, modern artillery such as tanks and armoured vehicles, we now have missiles that carry nuclear warheads and rockets that travel long distances. Drones and remotely piloted vehicles are creating new ways in which we can inflict harm on each other—they have the potential to drastically change the way we fight. For example in January 2020 Iranian commander Qasem Soleimani was killed by an American drone. The ever-increasing reliance on information systems is creating vulnerabilities to hackers and those hostile powers and terrorists who wish to disrupt our IT systems. For example in early 2020 Travelex customers were warned to keep a close eye on their accounts after hackers stole data. Cyber warfare is an evolving new form of conflict. But the influence of the military is not all negative. Military engineering and our collective spirit of adventure and thirst to probe new boundaries is driving the development of civilian rockets. We are contemplating further exploration of the ‘final frontier’—space.

War Warfare is a last resort. But if you must fight then you need better tools than your opponents—that is why engineering is important for the defence industry and the armed services. Like the rest of us, engineers have to reconcile three factors about warfare. First, that taking of human life is wrong. Second, that individuals, groups, cities and nations have a duty of care to defend their citizens and defend justice. Third, that sometimes protecting innocent human life and defending important moral values requires the use of force and violence. The concept of a ‘just war’—bellum justum—is pertinent after the brutalities of WWI and WWII—and usually divides into two parts. The first is jus ad bellum or justice of war or under what conditions is it justifiable to go to war. The second is jus in bello or what are the moral constraints on how a war is to be fought. These two parts are not unrelated although absolute pacifists will claim there are no moral grounds at all for war—even in principle. The science and engineering of nuclear weapons of mass destruction and instruments of torture graphically illustrates the ethical duty of engineers. War seems to be as old as mankind. One of the earliest civilizations in the world, Mesopotamia, was in a state of almost constant strife. Two of the oldest cities in the world were Jericho in Palestine and Uruk in Mesopotamia. Evidence of engineering work for defence and warfare is strong. Archaeologists think that a fortified city stood on the site of Jericho before 7,000 BCE. The walls of the fortress were 3

War

135

m thick and 3.9 m high and surrounded by a moat over 9 m wide and 3 m deep. Mesopotamians used a simple bow and arrow as early as 10,000 BCE. Sargon the Great of Akkad (2,334–2,279 BCE) unified Mesopotamia but rebellions had to be put down or invaders fended off. The Early Dynastic Period of Egypt (3,150–2,686 BCE) may have come about because the south of Egypt conquered the north. The State of Qin defeated the other states to unify China under Emperor Shi Huangti. This same pattern was repeated for other nations. For example, Philip II of Macedon (382–336 BCE) united the city-states of Greece. Over the centuries our propensity to go to war for all sorts of reasons such as gaining territory or power seems unabated. But one thing has changed—our ability to do more harm to each other through ever more powerful weapons. Weapons of mass destruction are a game changer. The future of warfare has become uncertain. Nuclear weapons threaten so much devastation that only a fanatic would now use one—but who would rule that out? It is a known unknown of the maximal kind. Terrorism has always been present throughout history, but the modern kind seems to be an unintended consequence of the breakdown of the cold war and of the previous imposition of dictatorships. There has been no peace dividend as forecast—rather the world has perhaps a wider cast of players, than ever before. Barbaric acts of terrorism such as the planting of improvised explosive devices, driving a car into crowds of people on the one side and remote killing by drones with lethal speed and quickness on the other tend to lead to an asymmetrical fear that breeds risk taking, the stretching of moral boundaries, a loosening of institutional controls, and cyberattacks from extremists. There is little sign of change as political ‘populism’ holds sway in 2019. In the longer term, an optimist might speculate that the uncertainty and potential consequences of the massive destruction of modern warfare, coupled with the increased levels of interaction between populations of countries through travel and social media, might help us better appreciate our interdependence—at all levels from individuals to entire nation blocs. Might these very same factors reduce our fear of others with their very different ways of life and values? Will they reduce our need to posture power (we have bigger and better weapons than you) to make them fearful of us? Might we realise that holding stocks of weapons to ‘balance’ what others hold is a situation of unstable equilibrium where a small perturbation—such as a small operational mistake—could lead to disaster? A simple mistake that could be the trigger that bursts the balloon of Barry Turner’s incubating accident. Such changes in attitude may seem unrealistic but could emanate out of a collective appreciation of the ingenuity put into past and present warfare as we struggle to find the ingenuity required to deal with future unknown forms of conflict.

Weapons To understand the role of engineers in war, let us first look at weapons and later we’ll look at targets (chiefly our infrastructure lifelines). The first craft-based weapons had only a confined effect as they were used in close combat. Over the centuries, primitive

136

7 Fighting

clubs, axe and spears have given way to swords and shields, then to the more widely felt battering rams, catapults, longbows and crossbows, and gunpowder in cannons, guns, bombs and mines. As science has developed so has our capacity to engineer more powerful weapons to kill greater numbers of people in one strike. In modern warfare, we use tanks, torpedoes, radar, sonar, nuclear and chemical weapons, drones and missiles controlled by computers and GPS. In the twenty-first century, warfare can sometimes seem more like a computer game with large-scale consequences— for example, when remote autonomous drones target individuals or weapons of mass destruction threatened to kill large numbers of people. Of course, WMD such as phosgene, chlorine and mustard gas were used in WWI. After WWII an early form of chemotherapy was based on a derivative of mustard gas. Chemicals such as Sarin (used in Iraq in 1988, Japan in 1995 and Syria in 2013) and VX are deadly nerve agents that are products of chemical scientific research. In 2018 the nerve agent Novichok, made from two non-toxic chemicals and easily transportable, was used to attack a Russian and his daughter in Salisbury, England. Although international treaties to ban such weapons have been agreed, their safe containment is nevertheless a formidable engineering problem as well as a political one. Although weapons have obviously become more lethal—able potentially to kill more people—Trevor Dupuy1 an American military historian has argued that, excluding nuclear weapons, mortality rates have not greatly increased. His data shows that adjustments in the tactics of battle, and the mobility and dispersal (men per sq. km.) of forces have reduced casualties as a percentage of committed forces for both sides of a conflict. Nevertheless, as weapon systems become ever more complex, with built-in AI, then we have to be aware of the risks inherent in technology triumphalism because there will be unintended consequences that could be very serious in scale.

Trebuchet—Early Ingenuity One of the first signs of ingenious engineering was a siege engine, called a traction trebuchet that appeared in China around the fourth century BCE and in Byzantium in the late sixth century. The counterweight trebuchet was still being used in the twelfth century. A large mass hung on a short lever arm which was lifted high in the air. The mass was released, swung down on the lever which rotated and propelled the ammunition at the other longer end through the air. Machines capable of throwing huge stones of around 150 kg (Fig. 7.1) were built. Reports say that some were even used to throw dead horses into a besieged city to spread disease. The people who made these and other war machines were called ingeniators (engineers). You may recall from Chap. 3 that Ailnolth was one of the first in England. His tasks were many and various and included maintaining the royal palaces and Westminster Abbey, the Fleet gaol in London, Windsor Castle, Orford and Framlingham Castles in Suffolk and constructing a new wharf on the Thames. He probably began as a carpenter craftsman with some technical training in the making of large war engines such as siege machines.

Gunpowder

137

Fig. 7.1 Reconstruction of a trebuchet at Château des Baux, France. Image by Quistnix. CC BYSA-3.0 via Wikipedia Commons

Gunpowder Gunpowder weapons changed everything—but only gradually. In 960 AD, Zhao Kuangyin founded the Song Dynasty in China, and united the fragmented states of the south. Gunpowder was made for the first time using saltpetre, sulphur and carbon. The Arabs referred to saltpetre or potassium nitrate as ‘Snow from China’. Around 1,000 AD it was being used for firecrackers or attached to spears to burst on impact. By the twelfth century, the Chinese were using crude hand grenades and an early form of rocket and cannon. Roger Bacon in 1216 was the first European to describe the process of making gunpowder but he kept it secret by describing his recipe in code. In 1904 Lt. Col. Hime deciphered the code as seven parts of saltpetre, five of young hazelwood (charcoal) and five of sulphur. The first metal cannon was the pot-de-fer or ‘iron pot’. It was an iron bottle with a narrow neck loaded with gunpowder and an arrow-like bolt that was set off with a heated wire through a touch hole. The weapon was used by both sides during the Hundred Years’ War but was still relatively rare. By the sixteenth century, cannons were made in a number of lengths and diameters with the general rule that the longer the barrel, the longer the range. Some cannon were 3.0 m long weighing 9,000 kg and needed large amounts of gunpowder. Castles were not immediately obsolete because of the cannon—but their importance declined. New fortresses had

138

7 Fighting

lower and more robust towers and were characteristically star-shaped on plan. By the fifteenth century, field artillery was improved through wheeled gun carriages, refined gunpowder, standardised calibres and better projectiles—however, cannon were still very slow and heavy. The mortar was a thick-walled short barrel gun useful for sieges firing bombs over high walls. Aiming was still largely guesswork (with no aligning ‘sights’) but gunners started to control the range by measuring the angle of inclination. The Marquis de Vauban was a Marshal of France and the foremost military engineer of his age. He was born in 1633 and lived to be 73. He designed both fortifications and methods to break them and promoted the idea of defence in depth, i.e. the coordination of multiple security measures. His work is considered to be of such significance that 12 groups of his fortified buildings in France make up the UNESCO World Heritage Site of the Fortifications of Vauban.

Samuel Colt The oldest handgun has been dated as 1288. Handguns were developed from the fire lance which was a kind of spear made of a bamboo tube containing gunpowder and a projectile. Matchlocks, wheel locks, flintlocks, percussion locks and revolvers followed as people tried to find mechanisms for the firing a handheld weapon. In the USA, the Colt became synonymous with the wild west. Colt’s Patent Firearms Manufacturing Company was founded in 1855 and mass-produced revolvers. Colt’s first two business ventures were disappointing but in 1847 the Texas Rangers ordered 1,000 revolvers during the American war with Mexico. Then Colt’s factory in Hartford, Connecticut, USA, supplied firearms both to the North and the South during the American Civil War of 1861–65. When Colt died in 1862, he was one of the wealthiest men in America. The later 1872, 0.45 calibre (11-mm-diameter barrel) Colt peacemaker is considered by many as the ‘gun that won the west’—the most popular revolver of the late 1880s with power and reliability. Colt pioneered the use of interchangeable parts put together on an assembly line as well as advertising, product placement and mass marketing.

Baron William Armstrong Seventeenth-century English ships were equipped with cannon that fired 15 kg solid shot that could penetrate solid oak from 90 m away. The Royal Navy adopted the carronade in 1779 because it was a lighter, shorter, cast iron cannon suitable for short range and easier to handle. In the 1850s, the carronade was replaced by steel cannon developed by William George Armstrong. In 1887, Armstrong was the first engineer to enter the House of Lords. He invented and developed many forms of modern artillery and gave his gun patents to the government. Armstrong first trained as a lawyer but at heart he was technical. Early in his career he saw that much of the

Baron William Armstrong

139

available power of waterwheels was being wasted so he decided to design an engine powered by water. The engine was actually built but attracted little interest. Not to be daunted he developed a piston engine to drive a hydraulic crane. He claimed that his crane could unload ships faster and more cheaply than conventional cranes and four were installed on the quayside in Newcastle. The success encouraged Armstrong to set up a business to manufacture cranes and other hydraulic equipment. The new company soon received orders—cranes for the Edinburgh and Northern Railways and for Liverpool Docks, as well as for hydraulic machinery for dock gates in Grimsby. The company expanded and was producing around 100 cranes per year and branching out into bridge building. In 1851, Armstrong reinvented an original idea by Joseph Bramah in 1812 (but not developed) of a hydraulic accumulator—a device to store energy.2 This was a cast iron cylinder that had a plunger inside under a heavyweight. Water was pumped under the plunger to lift it, so the cylinder was filled with water under the pressure of the weight. When required some valves on outlet pipes from the cylinder were opened and the water, under high pressure, was piped to a machine to do some useful work—for example, to operate a crane. Armstrong knew that the army had difficulty in manoeuvring heavy field guns. He decided to design a lighter, more mobile field gun, with greater range and accuracy. The result was a breech-loading gun that fired a shell, rather than a ball, from a barrel of wrought iron wrapped around a rifled steel inner lining. A rifled barrel has spiralling lines cast inside to make the flight of the shell more stable and accurate. In 1855, he had a five-pounder ready for inspection by a government committee. The gun was successful in trials, but the committee thought a higher calibre gun was needed, so Armstrong built an 18-pounder on the same design. The gun was superior to all of its rivals and Armstrong gave his patent to the British government. Armstrong formed a separate company Elswick Ordnance, in which he had no financial stake, to manufacture armaments only for the British government. However, there was opposition to the gun particularly from Joseph Whitworth of Manchester. Stories were put about that the new gun was too difficult to use, expensive, dangerous and frequently needing repair. Armstrong found the constant criticism very depressing and in 1862 the government stopped ordering the new gun and Elswick had no business. Compensation was agreed but the government would not release the company from its agreement not to sell armaments abroad. Eventually, the restriction was relaxed, and the company sold guns to both sides in the American Civil War. In 1864 the two companies, W.G. Armstrong & Company and Elswick Ordnance Company, merged to form Sir W.G. Armstrong & Company. Armstrong resigned from the War Office and decided to focus on naval guns. In 1868, the gunboat H M S Staunch was launched. The eighteenth-century bridge at Newcastle prevented ships reaching the Elswick works so Armstrong paid for a new swing bridge to be built. In 1882 Armstrong’s company merged with shipbuilder Mitchells to form Sir William Armstrong, Mitchell and Co. Ltd. specialising in building warships. The first vessels were for the Austro-Hungarian Navy and the first battleship was launched in 1887. Japan bought several cruisers some of which defeated the Russian fleet at the Battle of Tsushima in 1905.

140

7 Fighting

Up to the time of World War I artillery continued to evolve so that guns of any size could be moved quickly—there was a trend to greater mobility and smaller weapons. The German super-heavy guns in World War II ran on rails or caterpillar tracks. Cannon in the twentieth century were howitzers (short barrels, small charges propelling high trajectories), mortars, guns and autocannon (automatic rapid firing of shells). NATO defines the role of artillery as ‘the application of fire, coordinated with the manoeuvre of forces to destroy, neutralise, or suppress the enemy’. In modern warfare, the man-portable mortar is used for small highly mobile artillery, gunhowitzers are middle range and rocket artillery, or aircraft are long range. The internal combustion engine changed military engineering. The car and the aeroplane were specialised into important weapons. Military engineers had a major new role in supporting them. Tanks were built and used for the first time during World War I—but they were slow and unreliable. Stalemate in the trenches meant that a weapon was needed to cross the ‘no-man’ zones. Tanks protected against bullets and shell splinters and drove over barbed wire. Early tanks were mounted with machine guns and light artillery. Military engineers were gaining vast knowledge and experience in explosives so one of their roles was to plant bombs, landmines and dynamite. Tanks were used widely in World War II but bigger was not necessarily better. Even a slight thickening of the armour plate greatly increased the total weight. A bigger engine, transmission and suspension were required and so they became less manoeuvrable and hence more vulnerable. A balance had to be struck. Engineers began to design them for reliability, firepower, mobility and protection. As a result, targeting and ranging (fire control), gun stability, communications and comfort of the crew have all improved. Composite armour is being developed using different materials such as metals, plastic and ceramics—they are lighter but more expensive. Modern military engineers have three main jobs—engineering combat, supporting strategy and ancillary services. Engineering offensive combat is front line on the battlefield, for example, digging trenches and building temporary facilities, clearing improvised explosives, demolishing enemy bridges and roads and communications and rescuing damaged vehicles such as tanks. Supporting defensive strategy is communications, airfields, ports, roads, bridges and railways communication. Ancillary services are, for example, mapping and disposing unexploded bombs. In peacetime, military engineers are preparing for war situations and logistics and often have to respond to man-made and natural disasters.

Radar Modern radar is important in transport systems and in engineering combat. The first simple elements of a radar system (RAdio Detection And Ranging equipment) were built by Christian Hülsmeyer in Germany in 1904. He realised that the work by Clerk Maxwell and Hertz on electromagnetic radiation could be exploited. His apparatus called a ‘Telemobiloscope’ used a spark gap transmitter and could detect a distant

Radar

141

object but not how far away it was. At first his work was presented as an anti-shipcolliding system. Scotsman Robert Watson-Watt was a descendent of James Watt. After graduating in 1912, he researched radio waves for a while before joining the Meteorological Office—they were interested in his ideas for detecting thunderstorms using radio waves. By 1935, following a suggestion by a young colleague, Arnold Wilkins, and working together they demonstrated that they could detect an aeroplane and Watson-Watt was granted a patent for radar. Much later at a conference in 1953, he is reported to have said to Hülsmeyer ‘I am the father of radar, whereas you are its grandfather’.3 In the period up to WWII, a number of countries developed radar systems. Now radar is used widely for a whole range of applications ranging from shipping to weather forecasting. A radar transmitter sends out radio waves or microwaves to detect objects such as aircraft, ships, spacecraft, guided missiles, motor vehicles, weather formations and terrain. The waves reflect back from the object and are picked up by a receiving antenna. The signals are then processed to find the range, angle and velocity of the object. Radar had a decisive impact on the outcome of WWII as did, perhaps rather surprisingly, polythene. Polythene was an important insulating material for radar cables and a closely guarded secret during WWII. The material enabled the Allies to gain advantage in long-distance air warfare, particularly in the Battle of the Atlantic so large quantities of this new substance became important for the war effort. Reginald Gibson and Eric Fawcett working for Imperial Chemical Industries (ICI) in 1933 created it by accident. It was an unintended consequence of a study of the effect of high pressure on chemical reactions. Their equipment was crude and considered unsafe, so experiments were suspended. In 1935, Michael Perrin and colleagues restarted the work in the same laboratory and took out a patent in 1936. The first commercial manufacturing plants came in 1939. In 2018, annual global production is around 80 million tonnes for plastic bags and packaging and bottles. The principle of the Unintended is alive and well as plastic waste pollutes our oceans. The vice-president of the EU Commission, Frans Timmermans, is reported to have said in 20184 that Brussels’ priority was to clamp down on ‘single-use plastics that take five seconds to produce, you use it for five minutes and it takes 500 years to break down again….. 50 years down the road we will have more plastic than fish in the oceans…..If children knew what the effects are of using single-use plastic straws for drinking sodas……they might reconsider and use paper straws or no straws at all’.

Nuclear Weapons WWII was, of course, raging when atomic bombs were developed, tested and used in 1945—what happened in Japan was so awful that fortunately since then no nuclear weapons have been used that way. The science behind the bombs was developed during the first part of the twentieth century. The temptation to develop and use them to end the war was too great. USA, UK and Canada collaborated on the Manhattan

142

7 Fighting

Project during WWII. The American engineer who led the U.S. Office of Scientific Research and oversaw the Manhattan Project was Vannevar Bush. The United States Army Corps of Engineers officer Lieutenant General Leslie Richard Groves Jr oversaw the construction of the Pentagon and directed the Manhattan Project. In the 1950s and 1960s, the USA and the USSR vied for nuclear supremacy with the massive political and cultural problems of the Cold War. The first early warning systems were developed because, through the use of long-range rocket missiles, the warning of a nuclear attack had become a matter of hours or even minutes. In 2018, new posturing by nations claiming to have bigger and better weapons than their foes are creating new dangers for world peace. It is perhaps too idealistic to expect engineers to oppose such bravado—it is difficult to imagine that world leaders would actually press the nuclear ‘button’—that is one scenario that would trump the effects of climate change. There is always a chance that new talks might alleviate a tense situation. After the fall of the USSR, the proliferation of nuclear weapons became the big issue. India’s first test explosion was in 1974. The Indian test resulted in Pakistan successfully exploding devices in 1998. The former Soviet bloc countries of Belarus, Ukraine and Kazakhstan returned their warheads to Russia. South Africa dismantled its nuclear weapon programme in the 1990s. Israel is thought to have nuclear warheads. North Korea has tested weapons as recently as 2017. Many fear that despite denials Iran has a nuclear weapons development programme but in 2016 the International Atomic Energy Agency has verified that Iran is doing what is necessary to ensure their nuclear programme is and remains peaceful. In 2020 tensions in the Middle East have cast new doubts on Iranian nuclear policies. Clearly, one of the major issues for a peaceful world is that the enormous destructive potential of nuclear weapons is controlled. A fundamental principle of the nuclear safeguarding is that the verification is independent of the country and is performed by international inspectorates. The Treaty on the Non-Proliferation of Nuclear Weapons tries to ensure that nuclear weapons are not transferred to any non-nuclear weapons states. More than 180 non-nuclear weapons states are now party to the treaty. In other words, they have agreed that the International Atomic Energy Agency (IAEA) must apply safeguards on all their nuclear material and that the IAEA’s ability to detect undeclared nuclear activities needs to be continually developed. The European Union applies the safeguards to ensure that nuclear materials in the EU are not diverted away from peaceful use. After the Gulf War in the 1990s, countries began to realise that nuclear safeguarding needed to be strengthened. Iraq had carried out a nuclear weapons programme despite being part of the agreement not to do so. Engineers have an important role in controlling the proliferation of nuclear weapons. Nuclear safeguards are about verifying that civil nuclear materials are properly accounted for and are not diverted to undeclared uses as well as containment and surveillance. New systems of collaboration and verification are urgently needed.

Rockets and Missiles

143

Rockets and Missiles The engineering of rockets and missiles were a precursor to the exploration of space. A Russian Konstantin Tsiolkovsky pioneered rocketry and astronautics and the idea of reaching outer space. In 1903, he calculated that a multistage rocket fuelled by liquid oxygen and hydrogen was possible. In 1915, American Robert H. Goddard did some experiments with solid-propellant rockets (Fig. 7.2). He tried various solid fuels and measured the velocities of the burning gases. He flew a liquid-propellant rocket in 1926 but only reaching 56 m over a couple of seconds. The following year Slovak Aurel Stodola published a reference book for jet propulsion engineers. Hermann Oberth published a book in 1923 about rocket travel into outer space. His writings were important because they triggered many small rocket societies all over the world. One society, the Verein fur Raumschiffahrt (Society for Space Travel), led to the development of the V-2 rocket by Wernher von Braun used in WWII. Following the war von Braun worked in the USA for NASA. Many dub him as the ‘father’ of rocket science. Solid-fuel rocket engines are giant fireworks. Although they are very powerful, they cannot be switched off or controlled in any way, so they are typically used only during lift-off. Unlike aeroplane jet engines, which take in air as they fly through

Fig. 7.2 Robert H. Goddard standing with the first successful launch of a 3 m liquid fuelled rocket in 1926. Image by Esther C. Goddard. Public domain via Wikipedia Commons

144

7 Fighting

the sky, space rockets have to carry their own oxygen supplies (oxidizers) with them because there is no air in space. Liquid-fuel engines pump liquid hydrogen (the fuel) and liquid oxygen (the oxidizer) into a combustion chamber at the bottom of the rocket, burn the mixture and the hot gases fire backward from rocket to propel the rocket forwards. Most of the fuel is used in the first few minutes of a space mission to achieve an escape velocity from Earth of around 7 miles per second. The oxygen and hydrogen burn at a very high temperature, which makes the engine more efficient and powerful. However, before combustion, both substances are stored at extremely low temperatures to keep them liquid. More fuel can be stored as liquids than as gases. Also, liquid-fuel engines can be switched on and off during flight. The USA National Aeronautics and Space Administration (NASA) was created in 1958 because the West and the USSR were in a Cold War ‘space race’. In 1957, the USSR launched Sputnik 1, the world’s first artificial satellite. The USA launched Explorer 1 in 1958. What followed is well documented. First Russian Yuri Gagarin completed an orbit of the Earth in 1961. The USA Project Mercury single astronaut programme (1961–1963) proved that a human could survive in space. Project Gemini (1965–1966) had two astronauts practicing space operations such as rendezvous and docking of spacecraft, followed by Project Apollo (1968–1972) to explore the Moon. Missions to the Moon (Ranger, Surveyor and Lunar Orbiter), Venus (Pioneer Venus), Mars (Mariner 4, Viking 1 and 2) and the outer planets (Pioneer 10 and 11, Voyager 1 and 2) followed. There was extensive research to improve safety, reliability, efficiency and speed. Remote-sensing Earth satellites for information gathering were built and launched as were satellites for communications and weather monitoring and an orbital workshop for astronauts called Skylab and a reusable spacecraft for travelling to and from Earth—the Space Shuttle. Space stations started operating in 1969 when two Russian Soyuz vehicles were linked in space. The construction of the International Space Station (ISS) began in 1998—helped by the first reusable spacecraft the Shuttle. Until recently the ISS was only available for governments but new opportunities now exist for commercial and academic use facilitated by the Centre for the Advancement of Science in Space (CASIS).5 The International Space Station took 10 years and over 30 missions to assemble and is the result of unprecedented scientific and engineering collaboration among five space agencies representing 15 countries. The space station is approximately the size of a football field, an incredible 460 tonne, and permanently crewed platform orbiting 240 miles above Earth, and about four times as large as the Russian space station Mir and five times as large as the U.S. Skylab. In February 2018, SpaceX’s privately financed Falcon Heavy rocket was the first to take off and land with reusable components. Next, they will test the repeatability of the operation and if successful launch costs will be dramatically reduced. SpaceX and other private companies are ushering a new era of space travel with even bigger and better rockets are planned over the few years–promising exciting new developments ahead such as space tourism (for the very rich) and travel beyond the moon. At the same time, large 3D printers will be used to produce cheaper smaller rockets to launch small satellites.

Rockets and Missiles

145

SpaceX was founded in 2002 by Elon Musk, a South African-born businessman, investor and engineer. After graduating in physics and economics, he founded a software company which he sold at a considerable profit after only 4 years. He then founded the company that was to become Paypal which he also sold at a considerable profit. As well as SpaceX where he is CEO as well as chief technology officer, he also co-founded Tesla and became CEO whilst they built the first electric sports car, he co-founded SolarCity to provide solar power and has tested the feasibility of building a hyperloop.6

Targets Military bases are major targets in war. All through the ages, networks of military bases have been used to deploy forces. Examples are Roman forts (Hadrian’s wall), Norman castles (Kenilworth), sixteenth-century arsenals (Kremlin), fortifications (Maginot) infantry barracks and garrison towns (Catterick), naval dockyards (Pearl Harbour) and modern hubs like Camp Bastion in Afghanistan. Communications networks are targets too including roads, bridges and telecommunications. Civilian targets are legally (if not practically) protected in war—Protocol I to the Geneva Convention was drafted to protect them in a ‘just war’. To be legal an attack has to be a military necessity—on a military target with any harm to civilian property proportional. A war crime is when an attack on civilians is intentional. The evidence of engineering work for the defence of Jericho before 7,000 BCE is also evident that communities and states organised armies and advanced into other territories. Natural barriers such as rivers and mountains were defences. Apparently, the army of King Croesus (560–547 BCE) of Lydia (part of modern Turkey) couldn’t cross the Halys River (now known as the Kizilirmak or Red River in modern Turkey). Thales of Miletus was one of the soldiers and he organised a team to dig upstream and form two shallower channels that could both be waded across—quite a major engineering effort. Bronze or Iron Age forts were built on high ground to defend against warring tribes and invaders. Hadrian’s Wall was built by the Romans in AD122-232 and stretched 73 miles, coast to coast with military forts at 5 mile intervals along its length. Some hill forts were abandoned during the Roman occupation but reoccupied afterwards to defend against Anglo-Saxon invaders and then the Anglo-Saxons used them against the Viking invaders. The first timber castles were built by the Normans because after 1066 they needed to protect their new kingdom. However, they were not as strong as required so from about 1100 castles were built in stone. The White Tower at the Tower of London was started in 1070. Concentric castles were castles inside castles. They therefore had two or three walls around the keep and the inside walls were built higher than the outside walls to make defending easier. A deep wide ditch called a moat, often 10 m deep and 4 m wide, surrounded the whole castle. The age of castles lasted for nearly 500 years but became less effective as cannons became available to knock down walls.

146

7 Fighting

Ancient military harbours were often inside merchant harbours, protected by an outer wall, and places where ships could be beached, or cargos transferred by water. For example, Ostia was the harbour city and naval base for ancient Rome on the mouth of the River Tiber. King John of England had a dock built at Portsmouth in the early 1200s. Possibly the first dry dock was built there for Henry VII in 1495. The carrack, Mary Rose was rebuilt there in 1536, subsequently capsized in 1545, was raised in 1982 and is now on display in a purpose-built museum. The modern engineering challenges of building ports are many—examples are installing piles and retaining walls, underwater structures including inspection and maintenance, breakwaters, lock gates, dredging channels, managing siltation, dockside services including terminal buildings, handling containers and the effects of rising sea levels. An example of a special harbour engineering project was the construction of the Mulberry harbours in 1944. They were temporary portable docks built by the British for the Allied invasion of Normandy (Fig. 7.3). After holding beachheads following D-Day, two prefabricated harbours of reinforced concrete were floated in sections across the English Channel. The reinforced concrete caissons were sunk in place and used to create breakwaters. Roadways bridged between moored pontoons and pier heads where the ships were unloaded. The pier heads had four legs resting on the seabed but floated with the tide. Allan Beckett was the British engineer who designed the floating roadways and anchor systems. He was raised in East London

Fig. 7.3 Omaha Mulberry harbour. Public domain via Wikipedia Commons

Targets

147

and graduated in Engineering. At the outbreak of war in 1939, he volunteered for military service and was assigned to the Royal Engineers. Apparently in 1942 Winston Churchill sent a terse memorandum to the Chief of Combined Operations, Lord Louis Mountbatten, headed ‘Piers for use on beaches’. ‘They must float up and down with the tides’. ‘The anchor problem must be mastered. Let me have the best solution worked out. Don’t argue the matter. The difficulties will argue for themselves’. Eventually, Beckett was given a sketch marked ‘Top Secret’. The sketch showed a series of pontoons, a mile long, linked by bridges which covered a stretch of water that was shallow at one end and deep at the other. The caption read simply ‘Piers for flat beaches’, with no explanation of what they might be intended for. Beckett made a tin-plate model of his proposed floating roadway, with lozenge-shaped bridge spans joined with spherical bearings. Of the three designs submitted for testing, only Beckett’s bridge survived rough weather. The Mulberry harbours were used for 10 months—over 2.5 million men, 500,000 vehicles and 4 million tonnes of supplies were landed. Later that year the harbours were severely damaged by storms and abandoned. Beckett went on to be a senior partner of Sir Bruce White and partners consulting engineers. Air raid or bomb shelters were widely used in WWII as were some London Underground stations. Fallout shelters were built during the Cold War to protect against radioactive debris. The engineering of the protection of modern military installations like Camp Bastion was complex. High security perimeter fencing was supported by tight security procedures and various sensors and radar scanning to detect intruders and movement. In the twenty-first century, large-scale fortifications have largely become obsolete and replaced by digital detection systems. Tracking of friendly and hostile targets by radar together with drones, Unmanned Aerial Vehicles (UAVs), aircraft and satellite reconnaissance all combine to provide ‘situational awareness’ of possible threats.

Defence Nations have to defend themselves against aggressors and so they either buy in weapons or make them through their defence industries with big budgets long-term planning horizons and consequent vulnerabilities to rapid change. Very large complex equipment such as tanks, fighter jets, ships and aircraft carriers take a long time to plan and make and budgets are often considerably overreached. The fall of the Berlin Wall in 1989 was welcomed by all as the trigger for a peace dividend. Zandee7 refers to a dinner meeting in 1993 between the Deputy Secretary of US Defence and all of the captains of the American defence industries—which, he says, is alluded to as the last supper. As a result, the big five US defence companies consolidated as later did the European companies. Zandee argues that process has gone far enough because we need credible forces in an uncertain and unstable world. The problem, he says, for the fragmented European defence industries is not economic but political

148

7 Fighting

and governmental. As long as the majority of defence acquisition remains nationally planned, nationally ordered and delivered, the chances of creating a European Defence and Technological Base remain low.8 Out of this need has grown an appreciation by many of the engineers involved that an integrated systems engineering is required—but this understanding is not widespread amongst politicians or the electorate. Under current assumptions, weapons may have a short shelf life9 because powerful, disruptive forces will alter the defence industry over the next decades. Planners will continue to imagine future possibilities or scenarios with a rigorous treatment of uncertainties. But the consequences of unknown unknowns are potentially ‘end games’ if they involve nuclear weapons in the hands of people who are prepared to die for their cause. AI developments for military applications will be similar to those for civilians because the considerable expertise is held by civilian IT specialists. Warmongers will adapt civilian products to control weapons. After early experiences, responsible governments may yet decide that civilian drones are not yet reliable enough for war as they can’t be properly controlled onto sure targets without collateral damage—but individual terrorists will not be so scrupulous. There could be a shift from military to civilian control, from elected governments to unelected terror groups in war. The problems in this kind of shifting control lie deep in the technology. AI in robotics is being developed to do something similar to human decision making. But there are palpable differences. A computer senses the world around and then processes the incoming information through algorithms, with a choice of actions. Autonomous systems build a model of their world and continually update it through cameras, microphones and possibly tactile transducers, and the data is ‘interpreted’ to create an effective model that can be used to make decisions. The dependability of the world model and updates are key issues. Autonomous unmanned aerial vehicles (UAV) are relatively straightforward, because they rely on route maps, information about obstacles and no-fly zones. The data can be augmented in real time by radar and GPS coordinates. Civilian robo-delivery vehicles are currently being trialled in Arizona. Driverless vehicles such as cars or trucks carrying weapons are much more difficult. They need GPS maps but must also deal with other vehicles, pedestrians and cyclists, and anticipate what they may do. Driverless cars and some drones use combinations of sensors like LIDAR (Light Detection And Ranging), traditional radars and stereoscopic computer vision. But what we humans find easy takes a lot of computational power. A driverless vehicle or drone as a military weapon will make best guesses based on uncertainty distributions, and the potential unintended consequences are dire. In this chapter, we have seen that, whatever your ethical stance and like it or not, engineers created ever better ways for us to kill each other. The evidence is the ancient trebuchet, the log bow, clubs, swords, gunpowder, guns, cannon, modern artillery and nowadays missiles drones and rockets that travel long distances. Our ever-increasing reliance on information systems is creating new forms of warfare as hackers and hostile powers and terrorists attempt to infiltrate or disrupt our IT systems. But the influence of the military is not all negative. Military engineering and our collective spirit of adventure and thirst to probe new boundaries is driving

Defence

149

the development of civilian rockets. We are contemplating further exploration of the ‘final frontier’—space. It is evident that we humans tend to collaborate more effectively when faced with a common enemy—perhaps that is now climate change and the need for sustainable systems?

Sustainability Rather than a common enemy, perhaps a more positive representation of our need is that of an impelling and compelling common purpose. But that would require a total change of mindset. The challenge is to replace our fear of others with some form of common fear or goal. People who have a different way of life, and sets of values and beliefs, are not our enemies to be threatened and possibly attacked. Our common twin threat—nuclear destruction or climate change—is what we need to fear. Somehow, we have to realise, regardless of race gender, ethnicity and religion, that climate change will be the boss. We ignore the boss at our peril. As the extreme weather events increase—regardless of what is causing them—and we collectively begin to realise the seriousness of our predicament then we may be able to focus peoples’ attention on what really matters—sustainability. Our capacity to quarrel, fight and hurt each other seems to have gone beyond the modern idea of the just war—bellum justum. It’s a big step to get from the negative fear and hatred that underpins war and terrorism to its polar opposite—the love and concern for each other and our natural world that should underpin sustainability. Sustainability is a capacity to endure. The principle of Part tells us that it has to be done collaboratively and inter-governmentally in order to maintain global peace. War is about short-term destruction, whereas sustainability is about long-term survival. This contrast leads us into our two final chapters—well-being and human flourishing. Our ability to interfere with natural systems has grown to the point where it threatens our very survival—we have changed the climate and our organisations are becoming so increasingly fragmented that we have seem to have lost any sense of our complete interdependence in a finite world. As Carl Sagan says, ‘there is no hint that help will come from elsewhere to save us from ourselves’. Without human interference, natural systems remain diverse and they thrive—though that does not imply that every individual living species will survive. Think of the dinosaurs—they became extinct through no fault of theirs. We face possible extinction or at least devastating loss of life by our own actions. Sustainable development requires all five of our principles—as James Crowden has shown to good effect. The principle of Part tells us we are part of what we are changing and part of its unintended consequences. If one of those is the loss of human life on a major scale as natural disaster become more intense and frequent, then we have only ourselves to blame. To mitigate the possibility, we have a duty of care to each other to heed the principle of Preparedness—and work hard at encouraging each other to collaborate and develop trust. That will require the principle of Ingenuity to find new ways to work together across national and religious boundaries, to meet the common challenges we face

150

7 Fighting

and more importantly to value working together through international organisations like the UN to learn as we go along. The principle of Learning together to improve the ways we balance local and global efforts to meet basic human needs without destroying or degrading the natural environment. Above all the role of the engineer has to change. Inwardly by helping engineers to see themselves as a people profession, and outwardly by helping society at large to appreciate much more the value that engineers bring to our lives. So, the engineering of sustainable human well-being is critical. End Notes 1. Dupuy, TN. (1980), The Evolution of Weapons and Warfare, Bobbs-Merrill, New York. 2. Gibson, JW, Pierce, MC. (2009) Remnants of Early Hydraulic Power Systems, 3rd Australasian Engineering Heritage Conference, University of Otago, Dunedin, New Zealan, Nov. See http://www.ipenz.org.nz/heritage/conference/ papers/Gibson_J.pdf (Last accessed February 2019). 3. Bauer, AO. (2005) Christian Hülsmeyer and about the early days of radar inventions a survey. See https://aobauer.home.xs4all.nl/Huelspart1def.pdf (Last accessed February 2019). 4. The Guardian 16th Jan 2018, EU declares war on plastic waste, See https:// www.theguardian.com/environment/2018/jan/16/eu-declares-war-on-plasticwaste-2030 (Last accessed February 2019). 5. In 2011, NASA chose the Center for the Advancement of Science in Space (CASIS) to be the sole manager of the International Space Station U.S. National Laboratory. The mission of CASIS is to maximize use of this unparalleled platform for innovation, which can benefit all humankind and inspire a new generation to look to the stars. 6. A hyperloop is a sealed tunnel or tube through which a train or pod may travel with reduced air resistance to travel at high speed. See https://en.wikipedia.org/ wiki/Hyperloop (Last accessed February 2019. Musk, Elon (2013). Hyperloop Alpha. SpaceX. See http://www.spacex.com/sites/spacex/files/hyperloop_alpha20130812.pdf (Last accessed February 2019). 7. Zandee, D. (2013) The future of European defence industry, Clingendael Institute, The Hague, The Netherlands. See https://www.clingendael.org/sites/ default/files/pdfs/The%20future%20of%20European%20defence%20industry. pdf (Last accessed February 2019). 8. European Parliament Directorate-General for External Policies, (2013) The development of a European Defence Technological and Industrial Base (EDTIB), Policy Department. See http://www.europarl.europa.eu/RegData/ etudes/etudes/join/2013/433838/EXPO-SEDE_ET%282013%29433838_EN. pdf (Last accessed February 2019). The Abstract states ‘In 2007 the EU member states inaugurated a European Defence Technological and Industrial Base strategy. The gradual integration of national DTIB should lead to self-sufficiency for security of supply – but on a European rather than national level………The joint political vision has lost contact with the individual political and industrial

Sustainability

151

reality of the growing export orientation of European suppliers. In addition, security of supply depends ever more on the influx of civilian and defence goods as well as raw materials from beyond Europe’s borders. The EDTIB is trapped between the national and global developments.’ An EDTIB will need to be integrated, less duplicative and more interdependent. 9. Nurkin, T. (2016) The Future of the Global Defence Industry Jane’s Information Group. See https://cdn.ihs.com/www/pdf/The-Future-of-the-Global-DefenceIndustry.pdf (Last accessed February 2019).

Chapter 8

Well-Being

Engineering ingenuity from centuries of acuity.

Robert Langer’s Story Dr. Robert Langer is an American chemical engineer, a pioneer of bioengineering and one of a small number of Institute Professors at the Massachusetts Institute of Technology. He has used nanoparticles to carry targeted chemotherapy drugs to inhibit cell growth in cancer patients. His team have refined implantable microchips and polymer gels that can gradually release drugs, such as insulin and vaccines. They have created new biomaterials and engineered tissue such as skin, muscle, bone and even entire organs that can be grown to help victims of serious accidents or diseases who suffer from missing tissue or non-functioning organs. How have engineers improved our well-being? Bioengineering is one important aspect of medical engineering but, as we will explore in this chapter, earlier developments in public health engineering, contributions to agriculture, providing potable drinking water and disposing of waste have been crucially important in helping us to live healthier and longer lives. We take for granted the flushing toilet, but it is scandalous that in too many parts of the world people and particularly children have little or no regular access to clean water. Joseph Bazalgette arguably did more for our well-being than any other person when he ‘cleaned up’ London in the nineteenth century. The story of how Ben’s heart pacemaker (Chap. 1) came into being is a perfect illustration of engineering ingenuity to improve individual human well-being. But we start with Robert Langer and bioengineering. This form of engineering grew out of chemical engineering at the interface between molecular sciences (chemistry, biology and medicine) and large-scale engineering. Robert Langer’s work is worthy of a Nobel Prize—but there isn’t one for engineering. Given that Nobel was © Springer Nature Switzerland AG 2020 D. Blockley, Creativity, Problem Solving, and Aesthetics in Engineering, https://doi.org/10.1007/978-3-030-38257-5_8

153

154

8 Well-Being

born into a family of engineers and he himself was a chemist, engineer and inventor, it is surprising he didn’t include engineering, technology or mathematics. To remedy this in 2013, the UK Royal Academy of Engineering set up the Queen Elizabeth Prize for Engineering. It is now the most valuable of all in terms of money—worth around £1 million and as illustrious as the Nobel Prize and the Fields Medal for Mathematics. In 2015, Robert Langer was awarded only the second ever Queen Elizabeth Prize for engineering. Robert Langer’s interest in chemistry began when his parents bought him a chemistry set. At the age of 11, he set up a small laboratory in the basement of his house in Albany, New York. He read chemical engineering at Cornell University and then obtained a doctorate in chemical engineering from MIT in 1974. When he graduated, he was offered jobs in the oil industry, but he decided that he wanted to use his engineering education to help people more directly. From 1974 to 1977, he worked as a postdoctoral fellow for cancer research with surgeon Dr. Judah Folkman at the Children’s Hospital Boston and at Harvard Medical School. There he developed new and innovative ways of restricting the growth of tumours by delivering a controlled release of drugs. Initially, his work was received with a great deal of scepticism—nine of his initial research grant applications were rejected—but he persevered. Eventually the FDA (Food and Drug Administration) approved his polymer-based treatment for brain cancer in 1996. As a result, many products of his research are now being used through a number of spin-off companies. He won a large array of prizes and awards, including the Charles Stark Draper Prize (2002), the 2008 Millennium Prize, the 2013 Wolf Prize and the 2014 Kyoto Prize. He was named as one of the 25 most important individuals in biotechnology by Forbes Magazine and CNN (1999) and Bio World (1990), and as one of the 100 most influential people in America by Time magazine (2001). He has changed the lives of hundreds of students ranging from medical doctors through electrical and chemical engineers to physicists and chemists. Many of them have been elected to the USA National Academy of Engineering and the UK Royal Academy of Engineering and many lead their own companies and research labs across the world. Professor Susan Margulies is another who engineers tissue—this time at Georgia Tech. After graduating in mechanical engineering in 1982, she did a Master’s degree and a Ph.D. She now works on preventing and treating injury. Human cells can stretch without damage to a limit, after which they are damaged. Her goal is to find the thresholds and the factors that influence them. She does experimental and theoretical work on head injury to identify mechanical properties on models of animals and instrumented dolls, as well as data from patients. Her research on lungs includes models of animals with pulmonary disease, and she studies lung function in vivo and in vitro.

Well-Being

155

Well-Being Robert Langer and Susan Margulies are engineers working to help improve the health and well-being of people in very specific ways. Well-being is, of course, not just about health but includes general aspects of the quality of our lives. Every New Year we wish it to each other as health, happiness, contentedness and prosperity. Today we can expect to live longer than any of our ancestors. According to the World Bank1 life expectancy at birth was 31 in 1900, 48 in 1950 and today it is over 70 years. Of course, the variation in these figures is large. For example, in Japan, today’s figure is 85 but only 35 in Chad. There are also well-known differences of gender, economic circumstances, lifestyle and occupation. Life expectancy also increases as we grow older because so many children still die before they are 5. In 2015, more than 16,000 children under 5 died every day because of a lack of simple and affordable interventions such as breastfeeding, vaccines, medication, clean water and sanitation. Public health engineering is closely aligned with social medicine, preventing disease, prolonging life, increasing sanitation and hygiene, promoting clean air as well as organising health services. Proper shelter, food and water supply are important as is the control of noxious agents from many different sources such as traffic, farming methods, fertilisers, toxic chemicals, waste and sewage disposal, drainage, and heating and ventilating systems. Early humans thought of death and disease as supernatural—the work of a god or demon—so magic, and religion played a large role. Epidemics such as plague and cholera were thought to be a sign of poor moral and spiritual condition—to be mediated through prayer and piety. Nevertheless, by the thirteenth and fourteenth centuries, lepers and those with communicable diseases were being quarantined. The word is derived from the Italian phrase quaranta giorni which means 40 days because that was the time that ships arriving in Venice from infected ports were required to sit at anchor before landing. The Black Death of the mid-fourteenth century prompted more sanitary controls and attention to water supplies, waste and sewage disposal, and the condition of food. Miasma or bad air and poisonous vapours were commonly thought as causes of disease right up until the nineteenth century. However, as the influence of scientific reasoning grew, so too did new advances. For example, in the eighteenth century, Frenchman René Laënnec had invented a simple stethoscope and Englishman Edward Jenner began to inoculate for smallpox. Public health and hygiene received more attention and the first hospitals appeared in the form we might recognise today. Sanitary surveys showed the link between communicable disease and filth in the environment. During the industrial revolution in England, as workers moved to the cities, their health and welfare deteriorated. In the nineteenth century, public health institutions began to be formed, and international sanitary conferences held, culminating in the World Health Organisation set up in 1948. The greatest medical advance of the nineteenth century was initiated by the work of Louis Pasteur—the ‘father’ of bacteriology. Germ theory, the idea that certain

156

8 Well-Being

diseases and infections are caused by minute living organisms, led to many improvements such as pasteurisation, immunisation and water purification. During the twentieth century, scientific understanding became the basis for public health policy although the ways of delivering were shaped by the social and political values of individual nation states. The role of the state in the treatment of individuals began to increase. Governments around the world have, in their various ways, taken on sanitation, collecting data, investigating health issues, carrying out laboratory tests, regulating the environment, administering vaccines and caring for the vulnerable. As a consequence, national spending on health has risen steadily in most countries. In recent years, the level of spending has become controversial because of increasing traffic air pollution, ageing populations, increased expectations, rising labour costs and new products of medical engineering. Medical engineering has resulted in significant new benefits but also appreciable new costs. Medical and surgical procedures, such as angioplasty (widening arteries or veins with stents), laparoscopy (keyhole surgery with cameras) and joint replacements, that stem from collaborations between doctors and engineers are now commonplace. New drugs such as biologic agents for treatments are made from living organisms. New medical devices such as MRI (Magnetic Resonance Imaging) and CT (Computerised Tomography) scanners, implantable pacemakers, ventilators and incubators are impressive but not cheap. New support IT systems, based on the new discipline of health informatics, collect, manage and share electronic medical records and transmit information. The problems are that they are complex, require substantial investment, and there have been controversial successes and failures. The USA Standish Group report that in 2015 only 29% of IT projects were successful, 19% failed and 52% were ‘challenged’—meaning not meeting all success criteria. Other reports quote between 40 and 70% projects failing. The USA spends more than $250 billion each year on IT applications development across all industrial sectors. We can’t cover here all of the many important factors that affect the well-being we enjoy today, such as exercising, eating well, getting enough sleep and enjoying good company. But we can look at how engineers have helped us to eat, drink and dispose of our waste safely—some of our most basic physiological needs for sanitation and well-being. Then we can follow in some detail the story of how Ben’s heart pacemaker came into being as a recent example of highly successful medical engineering.

Agricultural Making Farming is a form of environmental and public health engineering. We change our habitat to do something useful—produce food. The dream is to eat well, the reality is eating well. We clear vegetation, till the soil and domesticate plants and animals. We selectively breed organisms to develop characteristics that increase their usefulness to us. We have developed plants to provide larger seeds and bigger fruit, and cattle to provide more milk or better meat. The potential that drives the flow of change is the need to eat. The flows of change are the ways we cultivate plants and animals. The

Agricultural Making

157

form and structure of engineering in farming are derived from the methods, tools and equipment used to till and prepare the land, sow and reap the crops, and breed and rear the animals. The scandal of agriculture is that, in too many parts of the world, people go hungry when there is no narrowly technical reason why. New engineering will bring even more improvements, but famine is essentially man-made. Corruption, war, crime and terrorism hinder the sustainable development of nations. Leadership and changes in governance require the social ingenuity of engineering in its widest sense. It starts with the simple humanitarian message—famine is not inevitable. That message should frame our common goal and be the basis on which our future social ingenuity will materialise. The more we appreciate the considerable ingenuity of the past the more likely we will be able to find the ingenuity needed for the future. The earliest humans may have evolved around 2.8 million years ago. The first evidence of the use of fire and cooking maybe a million years ago and the first modern humans, Homo sapiens, some 200,000 years ago. But they did not begin to farm until about 15,000–10,000 years ago. It is interesting that farming does not seem to have developed because food was scarce or because people were deprived. Population levels seem to have risen after people started to produce more food. One of the earliest farming tools was a crude pointed bent stick or tree branch used to hoe or scratch the surface of the soil and form a tilth to sow seeds. Over 4,000 years ago, these hand-held hoes became simple ploughs. Eventually they were pulled by animals such as oxen, camels and even elephants. Open shallow furrows were made by pushing the soil away to either side. Our ancestors soon found that the more they tilled the soil the better the germination rate and the higher the quality of the crops. In Egypt, the land was tilled twice a year with a wooden plough drawn by an ox or an ass, once to break the ground, and then to cover the seed (Fig. 8.1). The seed was sown by a funnel on the plough. Crop yields were greatly increased by irrigation. The ancient shadoof used on the Nile is still used even today. A long pole was pivoted on a support with a weight at one end to lift a bucket full of water at the other. The crops were cut with a curved bladed sickle. The mixture of straw, chaff and grain was tossed in the wind and as the grain fell to the threshing floor it was collected in baskets and stored in silos, pits and granaries. Survival depended on a successful crop. One wealthy landlord in the 6th dynasty (2345–2181 BCE) is reported to have owned 1,000 cattle, 760 asses, 2,200 goats and 1,000 sheep. The Romans and Greeks used teams of oxen (Fig. 8.22 ), and later fitted their ploughs with wheels to make them more manoeuvrable. Their tools were still largely wooden, but they did use iron for the shares, to make effective cutting edges. The plough had two small ears to make a rut but no mouldboard to turn the soil over. Iron was very valuable, so ploughshare metal was converted into weapons in wartime. Seeds were sown by hand. They were covered with a harrow, which may have had iron teeth or may simply have been a thorn bush. A more complicated plough, fitted with a wheeled fore-carriage, may have been used in northern Italy as early as the first century AD. Traditional methods are still used even today (Fig. 8.3). The open-field system began around 800 AD and lasted into the twentieth century in some parts of the world. By the late tenth century in some parts of Western Europe, a

158

8 Well-Being Figure

Date

Title

Innovation

8.1.1

c.1200

Ancient

Tilling the

BCE

Egyptian

land rather

plough. Image

simply

in burial

produced

chamber of

better crops.

Sennedjem on the west bank of the Nile near Thebes. 8.1.2

Ancient

Roman plough

Pulled by

Roman

with oxen

teams of oxen. Later Roman ploughs had wheels and iron for shares (cutting edges)

8.1.3

2013

Traditional ploughing in India

8.1.4

1730

Rotherham

Wooden with

Swing Plough

iron coulter

by Joseph

(cutter) ahead

Foljambe

of the iron plated mouldboard (turns over the soil) and shares (cutters).

Fig. 8.1 A timeline of the plough

Image

Agricultural Making

159

Figure

Date

Title

Innovation

8.1.5

1855

John Fowler

A steam

(Image

traction engine. engine

1916)

dragging a plough attached to a wire rope wound on a drum mounted below the boiler.

8.1.6

1902

Englishman

Internal

Dan Albone

combustion

patented his

petrol engine

Ivel general

powering a

purpose

farm vehicle.

agricultural vehicle.

8.1.7

Mid-

An Allis-

Note the

twentieth Chalmers WD

absence of

tractor of 1948

‘roll-over’

in USA.

safety bars or

century

cab.

8.1.8

Early

A Norwegian

Two sets of

twenty

Kverneland

mouldboards

first

reversible

(left and

century

plough pulled

right) with a

by an

powerful

American John

hydraulic

Deere 6200

lifting

tractor.

system.

Fig. 8.1 (continued)

Image

160

8 Well-Being

wheeled asymmetrical ‘heavy’ plough with a rudimentary mouldboard and a coulter (cutting blade) to turn the soil over was being used. The expansion of farmland between the tenth and thirteenth centuries came to an end with the climatic, pestilent and military disasters of the fourteenth century. Very bad weather in 1314–1316 drastically reduced yields. Floods destroyed reclaimed land. Plague followed the famine, bringing untold suffering. The Black Death in 1347 is said to have killed one-third of the population of Europe. Outbreaks followed for the rest of the century. The Hundred Years’ War desolated much of France. Consequently, much of the land could not be cultivated because there weren’t enough people. Many of the settlements were deserted. Agriculture began to recover in the fifteenth century but at first there were no changes in methods. Feudalism still reigned—peasants forced to work the land for a feudal lord. However, foreign trade expanded, and merchants began to import goods and make huge profits. Capitalism emerged during the seventeenth century as merchants began to dominate producers. The practice of the waged worker heralded a change from merchant—making money from trade, to capitalist—wealth from the owning and controlling the means of production. Early English law had required every ploughman to make his own plough and he alone could use it. The design changed little until the mid-1600s when the Dutch acquired a superior Chinese plough. They used it successfully on wet, boggy soil in England. Joseph Foljambe from Rotherham improved it and built and patented a new plough in 1730 (Fig. 8.4) which became known as the Rotherham or Yorkshire swing plough. The plough was wooden, but the fittings and coulter were iron, and the mouldboard and ploughshare were iron plated. It was the first to be factory produced, light and easily worked with a pair of horses. In 1763, James Small in Scotland experimented by varying the curvatures of the mouldboard and produced a shape that would turn the soil more effectively. His ‘Scots Plough’ was the beginning of the plough we know today. In 1750, the English population was around 5.7 million—by 1850 it was 16.6 million. This period (approximately) of rapid growth has been called the British agricultural revolution. The reasons for it are controversial but it seems that the largely organic methods were slowly superseded by energy-intensive methods. By growing legumes and other crops, the amount of nitrogen in the soil increased sustainably. But then much later (nineteenth century) chemical fertilisers were introduced and sustainability was undermined—one of today’s grand challenges (Chap. 9). Productivity (amount of food per worker) increased during the agricultural revolution. Consequently, the proportion of the labour force working in agriculture declined and workers moved to towns and cities as the Industrial Revolution began. The reasons for the productivity improvements may well have been the development of agrarian capitalism with landowners, tenant farmers and labourers, introducing better farm management and efficiencies and new engineering. Steam power began to take over from the horse. A number of patents were issued between 1830 and 1850. In 1854, the Royal Agricultural Society of England (RASE) offered a prize of £500 for ‘the steam cultivator which shall in the most efficient manner turn over the soil and be an economical substitute for the plough or spade’.

Agricultural Making

161

The first contest was in Carlisle in 1855. John Fowler was rewarded with a medal for his strenuous endeavours (Figs. 8.1.5 and 8.1.6). The judges said, ‘Steam ploughing as such had attained a degree of excellence comparable in point of execution with the best horse work’.3 And in 1867, David Greig wrote ‘The steam engine stands on the headland and hauls the implement to and fro by means of a wire rope. All treading and compression of the soil and sub-soil associated with horse cultivation is thereby entirely avoided and the implement is driven at a much more rapid pace, throwing up the soil to a greater depth and in a loose state enabling it to derive full benefit from the influences of the atmosphere……The person who farms by steam has a powerful and untiring force at this disposal such that he can afford to wait until his land is in an exact state for working’.3 The internal combustion engine soon became the chief source of power. The first engines were stationary, but by the 1890s engines mounted on wheels were selfpropelled. The first wheeled tractor was built in the USA in 1892 and slowly took over from steam in the early 1900s. A tractor (from the Latin tractus meaning pulling) converts power via a gearbox, not to go fast like a racing car, but to pull a heavy load. The consequent lack of speed may be irritating when you are following one on a public road but when you do just remember you may be grateful for its pulling power if your car gets stuck in a ditch. Within a few years, several companies were manufacturing tractors in Germany, the United Kingdom and the United States of America. In the USA, the number rose from 600 in 1907 to almost 3,400,000 by 1950. Most early ploughs were simply pulled behind the tractor. Harry Ferguson developed a three-point linkage in 1928 (Fig. 8.1.7) that enabled tractors to carry a vast array of implements with better manoeuvrability, safety and traction. The linkage has two lower connection points and one upper point in a triangle connected by arms, to form a stable connection. Any piece of equipment, such as a plough, can be attached and lifted, moved side to side turned or rotated. Hydraulic power using pressurised fluid to drive machinery had been developed by William Armstrong in northeast England in the late nineteenth century. Ferguson used a hydraulic drive to raise and lower the lower arms of his linkage with a separate upper arm to control the draft (depth) of the tillage equipment. The networks of narrow pipes, cables and rams are the ‘muscles’ of the machine as the tractor’s engine pumps high pressure fluid through the network to create movement and to do the work. Agricultural engineering developed rapidly in the twentieth century. In the 1930s, horse-drawn and early tractor ploughs operated with widths between furrows of only around 150 mm, because of the limited power available. The seed would be in rows just wide enough for the horse’s foot to pass while walking through the crop. Early reversible ploughs had furrow widths fixed at around 400 mm, but modern equipment have variable furrow widths. Four-wheel drives and diesel power came in the 1950s and 1960s, leading to the enormous tractors we often see today. Modern ploughs are reversible, which simply means that each one has two sets of mouldboards—one turning left going in one direction and one turning right on the return. And they are big (Fig. 8.1.8). Their hydraulic systems enable the plough frame to swing in line with the tractor and be stable. Ploughing equipment has to be properly aligned and adjustable to cope with differing conditions safely.

162

8 Well-Being

The contribution of engineering to agriculture has, in the past, been largely concerned with easing and improving the use of manual labour through better machinery and chemicals to improve crop yields and deal with pests. One engineering grand challenge is to remedy the way we have interfered with the nitrogen cycle largely through chemical fertilisers. ‘Precision agriculture’ using sensors is one possibility (Chap. 9).

Drinking Water Access to clean water is essential for well-being. Towns and villages grew around sources of water such as springs, wells and, of course, rivers. Some of the earliest water wells were in China possibly 6,000 years ago. The Babylonians, Persians and Greeks used underground clay pipes for water supply and sanitation. Around 500 BC, Hippocrates invented a bag filter to trap sediments that caused bad tastes or odours, but there was no understanding of the health risks. The Romans built major aqueducts such as the Pont du Gard (Fig. 8.2) in France and installed plumbing in their houses and public buildings. After the fall of Rome, the sophistication of Roman plumbing was lost for many years.

Fig. 8.2 Pont du Gard

Drinking Water

163

There is evidence that by the middle ages in the UK people were well aware that not all water was safe to drink.4 Polluted water mostly in towns and villages would have been avoided as would water from marshy areas or places of standing water. However, if they knew the water was coming from a good source, they would not be afraid to drink it—contrary to popular myth that people only drank ale, beer and wine in medieval times—after all, the water was free. People did drink ale but not just because the water was so bad. The brews were much weaker than modern beers and provided much needed calories. Wine was a drink of the well-to-do since it was expensive, but innkeepers sold wine by the cupful for all to enjoy. Engineered improvements in water supply and dealing with waste were a long time coming. In London, around 1236, a system of pipes (called the Great Conduit) was built to take water from a large fresh spring at Tyburn to a pumping house at Cheapside, and this fed cisterns all over London. Wealthy city dwellers could have a private pipe connecting them to the conduit system, but it was expensive. Most people drew water from the nearest cistern or paid a ‘cob’ or water carrier to bring them a day’s supply in tubs carried on a yoke. There was a rapid extension of waterworks in the seventeenth and eighteenth centuries. Sir Hugh Myddleton oversaw a project to take freshwater from the River Lea near Ware in Hertfordshire to New River Head in London, between 1608 and 1613. The consequent New River Company became one of the largest private water companies of the time and supplied the City of London. It was clear that the rapidly growing population needed more water networks and the water needed to be treated. Francis Bacon in 1657 experimented with sand filtration but the first water filters of wool, sponge and charcoal appeared in the 1700s. The Chelsea Waterworks Company was established in 1723 ‘for the better supplying the city and liberties of Westminster and parts adjacent, with water’. The company built ponds and reservoirs in Chelsea and Pimlico using water from the River Thames. The turning point came with the realisation that water had to be treated for domestic supplies. The first municipal water treatment plant designed by Robert Thom was built in Scotland in 1804 with the water slowly filtering through sand. This method works because a biofilm forms in the top few millimetres of the sand. As the water passes through, foreign particles and soluble organic material are adsorbed (held in a thin film). Sand filtration is still used today. Thom’s water was distributed by horse and cart—then about three years later, he installed water pipes. In 1854, John Snow began to think that cholera was caused by water contaminated by sewage. He used chlorine to purify the water, and from this came water disinfection. In the late nineteenth century, national governments started to install municipal water filters (sand filters and chlorination), and to regulate public water. For example, by the early 1900s, the many privately owned water companies were taken into public ownership by the British government but major parts were again privatised in 1973. Nowadays, engineers attempt to configure the energy of water flowing under gravity to integrate all water needs—supplying drinking water, preventing or controlling flooding and treating wastewater. Unfortunately, different organisations are usually responsible for the various services. In a properly integrated system, engineers want

164

8 Well-Being

to allow water to flow under gravity wherever possible and pump it to water treatment works and our taps only where absolutely necessary. They seek to understand the natural energy of the water flows and then to re-engineer them. The flows occur in reasonably well-defined areas of land called drainage catchment basins. All rainwater, melting ice and snow in the basin flow on the surface and percolate through the earth as groundwater towards a single outflow to the sea or ocean. As the multiple streams, lakes and rivers merge into a single river estuary then the flow rate reduces naturally. This happens because energy is dissipated through the resistance of the ground and flow is retained in various forms—engineers call it capacitance or storage of the potential for flow. Lakes are natural capacitors and reservoirs and floodplains are engineered to create stores of potential energy. Water flow is also delayed in inductors such as long coiled pipes, waterwheels, turbines and inertia barriers such as vegetation. Flooding and water supply problems often occur where the management of these various aspects of water flow, in a given catchment to provide different services, are not properly integrated. For example, where water extraction for drinking causes levels of groundwater not acceptable to a river authority, or the dredging of a canal causes flooding downstream. Proper integration requires a substantial degree of collaboration which is sometimes hindered by organisational and political misalignments. Water companies that supply drinking water and treat wastewater are only allowed to extract a certain amount of water out of rivers and underground sources—to protect our environment. At many waterworks, water is stored in reservoirs before being treated to make sure there is always water available—for example, for about 90 days. A number of different engineering specialists work together to supply the water in your tap. Civil engineers build and maintain the infrastructure of the reservoirs, dams and storage of water. They install the large pipes that distribute the water to various locations, such as the treatment works, and they build the buildings in which the work is done. Chemical engineers are responsible for water quality. They design and supervise the detailed processes of removing large objects like logs of wood or dead animals. They also use biological and chemical processes to clean the water and make it potable. The biological processes often use microbes that decompose organic matter and the chemical processes may include algaecides to kill algae, chlorine dioxide as a steriliser and disinfectant and ozone to kill bacteria and viruses. Mechanical engineers design, install and maintain all of the pumps and many other pieces of mechanical plant, machinery and equipment. Pipes and channels are fitted with valves, flow meters, water meters and filtration plant. Much of the equipment is powered by electricity. Electrical engineers design, build, install, manage and maintain power stations to generate the electricity for the national grid. They provide electrical motors and generators and lots of other electrical equipment too. Electronic engineers provide the instrumentation used to monitor and control the flow of water. Finally, computer engineers design, install and maintain the computer systems that control the flow of information necessary to keep the water flowing.

Waste

165

Waste Human waste needn’t be. In nature, there is no waste; everything is used. Human waste is a rich, valuable, inexhaustible material—as the behaviour of any dung beetle shows. Apparently, they prefer human and chimpanzee faeces over the rest. We have yet to learn the lesson of nature that recycling dominates in the natural world. In manmade systems, valuable resources of matter and energy are lost in waste disposal. We can do more to reduce pollution, minimise waste, recycle and treat and recover materials and energy, whether chemically, biologically or through heat. Among the deities worshipped by ancient Romans was the goddess Cloacina. They believed that she was vital to their good health and sanitation. They placed their trust in her to look after their sewers and infrastructure of public works. She eventually became their goddess of purity and the protector of sexual intercourse in marriage and known as Venus Cloacina. Initially, Rome’s drainage ditches were open channels, but around 500 BCE enclosed sewers were being built. By 100 AD, Romans living in the capital had access to public lavatories and the sewers took away rainwater as well as the sewage. The Romans were not the first to build toilets and sewers. At Skara Brae in the Orkney Islands, stone huts were found in the 1930s that had been inhabited between 3,200 BCE and 2,200 BCE. A drainage system in the village may have included an early form of toilet. In Northwest India and Pakistan in 2,000 BCE, towns were built with networks of sewers flushed with water. In Egypt, around 1,200 BCE rich people used a container with sand—which was emptied by slaves. After the fall of Rome, it wasn’t until the twelfth century that, at Portchester Castle in southern England, Augustinian monks built stone chutes leading to the sea so that when the tide came in the sewage washed out to sea. In many castles, the toilet was a stone seat on the top of a vertical shaft cut into the thickness of the walls. The leaves of a plant called woolly mullein (verbascum thapsus) were used as toilet paper. The plant has emollient and astringent properties and employed to treat a variety of skin problems—its various names include ‘old man’s blanket’ and ‘indian rag weed’, and in the western United States of America it is sometimes referred to as ‘cowboy toilet paper’. In Tudor England, men called gong farmers (named after the Old English word gang which means ‘to go’) dug out and removed human excrement from privies and cesspits. Gong farmers were only allowed to work at night, so they were also known as night men and the waste they collected was called night soil. Today, cesspits are cleared using specialised tankers. Pail closets (pail privies) and water closets became common by the nineteenth century. A pail closet was often a small outdoor room with a seat and a pail under. The pail was removed and emptied regularly but sometimes overflowed and was a health risk, to say the least. The more advanced water closet was used only in wealthy homes. Sir John Harrington was a godson of Queen Elizabeth I but, apparently, she banished him from court for telling risqué stories. In some disgrace, he went to live in a village called Kelston near Bath, England. He had a house built and to make life more congenial he devised the first flush toilet. He named his new invention Ajax, after the Greek

166

8 Well-Being

hero of strength and fearlessness. Ajax consisted of a pan with a pipe outlet in the bottom that was sealed with a leather valve. Harrington devised a system of handles, levers and weights to pour in water from a cistern and then to open the valve and let out the contents. Whether or not the Queen heard about his invention and wanted to try it out is unknown. But she did eventually forgive him, and she did visit Kelston in 1592. Naturally, Sir John showed off his new facility and the Queen, we are told, did use it. She was so impressed that she ordered one, but we have no record of her subsequent level of enthusiasm for her new convenience. Unfortunately, in spite of her enthusiasm the public remained faithful to the more familiar chamber pot. In those days, chamber pots were usually emptied from an upstairs window onto the street. In France, people below were warned by the cry ‘gardez-l’eau’, which some say is the origin of the modern English colloquialism for the toilet, ‘loo’. Sir John Harrington didn’t patent the Ajax. It was some two hundred years later in 1775 that the first patent for a flush water closet was awarded to Alexander Cumming (or Cummings) in London. Cumming was a Scottish watchmaker and instrument inventor. Little is known of his early life except that he was born in Edinburgh in 1733 and was apprenticed to an Edinburgh watchmaker. In the 1750s, he was employed by Archibald Campbell, 3rd Duke of Argyll at Inverary as an organ builder and clockmaker. He moved to England and by 1763 he had premises in Bond Street, London, and ‘had acquired a sufficient reputation to be appointed a member of the commission set up in that year to adjudicate on John Harrison’s timekeeper for discovering the longitude at sea’.5 In 1765, he made a clock for George III which also acted as a barometer and recorded air pressure against time—the first accurate recording barograph. In 1783, he was a founding Fellow of the Royal Society of Edinburgh. Cumming’s device was rather similar to the Ajax (Fig. 8.3). His idea was to use an S-shaped trap to retain water in the outlet pipe and hence stop sewer gases from entering the room. Today, we have lengthened his version and we call it a U bend, but the idea is essentially the same—a big step towards a healthier system. Two years after Cumming, Samuel Prosser patented a closet that had a plunger inside the bowl on top of the outflow pipe, and the user would lift it to release the contents. It wasn’t really very hygienic and was soon replaced by a design by Joseph Bramah. He used a hinged valve at the bottom of the pan. The result was successful enough to be used extensively on ships and boats (Fig. 8.4). Both Britain and the United States needed organisations to supervise and regulate public health. It wasn’t until 1848 that a Public Health Act ruled that every new house in Britain should have a ‘WC’ privy, ash-pit or cesspit. The Act established a General Board of Health to guide local authorities, but it had limited powers and money. It was the first step on the road to improved public health. Social reformer Edwin Chadwick argued that if the health of the poor was improved then fewer people would seek poor relief—money spent on public health would save money in the longer term. He proposed better drainage and sewers, removing refuse from houses, streets and roads, providing clean drinking water and appointing a medical officer for each town. In America, the Shattuck report of 1850 by the Massachusetts Sanitary Commission examined the serious health problems and grossly unsatisfactory living

Waste

167

Fig. 8.3 Alexander Cumming’s 1775 patent S-trap toilet. Public domain via Wikipedia Commons

Fig. 8.4 Bramah’s Water Closet 1778. Public domain via Wikipedia Commons

conditions in Boston. It recommended the setting up of a public health organisation with local boards of health in each town. In fact, the first one was set up in New York City in 1866. One of the first businesses to capitalise on these trends was set up by Thomas Crapper. His name is one that many associate with lavatories because it lives on as a

168

8 Well-Being

slang expression, but in fact the word ‘crap’ is of Middle English origin.6 He was a plumber, and while he did not invent the flush toilet, he did increase its popularity. He also made some important improvements, such as adding the ballcock valve. Crapper was a Yorkshireman who moved to London. He quickly took up and promoted sanitary plumbing. The company Thomas Crapper & Co. owned the world’s first bath, toilet and sink showroom, in King’s Road, London from 1870 until 1966. In the 1880 s, Prince Edward (later Edward VII) bought Sandringham House in Norfolk and asked Crapper to fit out his 30 lavatories with cedar wood seats. Further improvements on the construction of the ‘WC’ followed. In 1852, George Jennings patented a washout toilet pan—a shallow basin with a dished tray and water seal. The flush water drove the contents into the pan and then through the S-trap. Thomas Twyford, an English potter who took over his father’s business, revolutionised the water closet business in 1883 with the launch of the first one-piece closet made of china clay, rather than wood and metal. It was free standing, hygienic and cheap with a hinged wooden seat. The number of patents issued for water closets accelerated during the late nineteenth century—the potential market was huge. In America in 1870, John Randall Mann was granted a patent for a three-pipe closet using the suction of a siphon. Wash down toilets flush out the waste by free-flowing water. A siphon toilet has a smaller outlet so that some water is trapped. When there is enough water in the bowl a siphon pulls the contents out. In 1876, William Smith also patented a jet siphon closet. Thomas Kennedy, another American, improved on Mann’s design and patented a siphonic closet with only two delivery pipes—one flushed the rim and the other started the siphon. The U.S. Patent Office received applications for 350 new water closet designs between 1900 and 1932. Two of the first granted in the new decade were to Charles Neff and Robert Frame of Newport, Rhode Island. They produced a siphonic washdown closet in 1900 but which was still prone to overflowing. Fred Adee fixed the problem 10 years later by redesigning the pan, to stop the messy overflows and gave birth to mass production of the siphonic closet in America. By 1890, special toilet paper on rolls was on sale in the USA and by 1900, houses for workers had inside lavatories. Improving the construction of the toilet for hygiene within the house was only one part of the problem; managing the waste disposal was quite another. The year 1858 became known, in London, as the year of the Great Stink. Meetings in the Houses of Parliament were suspended because of the foul smells. The problem was not confined to London. Something had to be done.

Joseph Bazalgette Joseph Bazalgette was the game-changing engineer for public health well-being. In 1859, he began the task of designing and supervising the building of London’s sewers. It was to make him one of the most influential engineers of his time. Born in London in 1819, Bazalgette began his career as a railway engineer, and in 1842

Joseph Bazalgette

169

Fig. 8.5 Joseph Bazalgette. Public domain via Wikipedia Commons

he set up in private practice. The London Metropolitan Commission of Sewers was created in 1848 with Edwin Chadwick as one of the Commissioners, but was quickly succeeded, in 1855, by the Metropolitan Board of Works for London—the first organisation to supervise public works in a unified way over the whole city. That too was to prove short lived, replaced by the London County Council in 1889. Bazalgette was appointed assistant surveyor to the Commission in 1849 and took over as the ‘Engineer’ in 1852 and immediately began to plan what needed to be done. At the time, the Thames was an open sewer with no fish or other wildlife. Cholera first struck London in 1832. Over 14,100 people died in 1848–1849 and a third outbreak in 1853 killed 10,738 people. Around 600 more died in 1854. Cholera was widely thought to be caused by miasma foul air but one doctor, John Snow, suggested that cholera was spread by contaminated water. Few agreed with him. Nevertheless, in 1847 the Board ordered that all cesspits should be closed and that house drains should connect to sewers and empty into the Thames. The Thames continued to be highly polluted and following the Great Stink of 1858, Parliament passed an act to permit Bazalgette’s plans to go into operation in spite of the cost. Bazalgette’s scheme intercepted the flow of foul water from old sewers and underground rivers and diverted it along new low-level sewers. These were built behind embankments on the riverfront and took the flows to new treatment works. Eightytwo miles (132 km) of underground brick main sewers and 1,100 miles (1,800 km)

170

8 Well-Being

of street sewers were constructed. The scheme involved major pumping stations at Deptford (1864) and at Crossness (1865) on the Erith Marshes, both on the south side of the Thames, and at Abbey Mills (in the River Lea valley, 1868) and on the Chelsea Embankment (close to Grosvenor Bridge; 1875), north of the river. The system was opened by Edward, Prince of Wales in 1865, although the whole project was not actually completed for another ten years. Bazalgette was very thorough in his work. It is said that he checked every connection. The records contain thousands of linen tracings with handwritten comments in Indian ink on them, such as ‘Approved JWB’, and ‘I do not like 6 in. used here and 9 in. should be used JWB’. Bazalgette’s thinking is illustrated by the way he decided on the diameter of the sewers. He established where the population was most dense and then allocated a generous allowance of sewage production for each person. From that total he calculated a pipe diameter. But he must have thought to himself, we’re only going to do this once and what about things unforeseen—so he doubled the diameter. This foresight dealt with the increase in population—for example, when tower blocks were built in the 1960s. His first smaller pipe diameter would have overflowed by then, but because he doubled it, the system is still coping even now. Most importantly the new sewer system removed the cholera, and reduced typhus and typhoid too. The Albert and the Victoria Embankments were both open by 1870. They reclaimed ground for riverside roads and gardens behind their curved river walls and replaced the tidal mud of the Thames shore. The Victoria Embankment contained Bazalgette’s low-level sewer, as well as a service subway and the underground railway. The Chelsea Embankment reclaimed over 52 acres from the Thames and was completed in 1874. During all of this work Bazalgette trained young civil engineers and gave advice to other towns and cities, home and overseas, including Budapest, as well as countries such as Mauritius. So, what of Chadwick and Bazalgette’s legacy? For London and most of the western world, he profoundly improved our health. But not for many other people around the world who still, scandalously, lack access to clean water. As Rose George, writing in 2008 on toilets and human health in the early twenty-first century reflects ‘when…. human achievement seems limitless, 2.6 billion people lack the most basic thing that human dignity requires. Four in ten people in the world have no toilet. They must do their business instead on roadsides, in the bushes, wherever they can. Yet human faeces in water supplies contribute to one in ten of the world’s communicable diseases. A child dies from diarrhoea—usually brought on by faecal-contaminated food or water—every 15 s. Meanwhile, the western world luxuriates in flush toilets; in toilets that play music or can check blood pressure, where the flush is a thoughtless thing, and anything that can go down a sewer’.7 Again, as for food shortages around the world, there are no narrowly technical reasons for the situation Rose George describes. Yet again what is required is social ingenuity at all levels of society. As the rest of us begin to appreciate the creative, inspirational, uplifting practical ingenuity that has gone before we have a platform to start applying pressure onto world leaders to find ways for our common humanity to reassert itself. We need leaders who will facilitate and articulate a common vision, with zero tolerance of corruption, war and crime.

Joseph Bazalgette

171

The flush toilet was a good example of how previously unimagined engineered objects can come into being. Someone’s new dream through collective ingenuity becoming, over a number of years of evolutionary development, everyone’s new reality. Engineers are now putting their minds to developing hygienic, low water using toilets that are not connected to a sewer system. For example, Sabine Schober won the 2013 World Toilet Organisation Design Award in 2013 for a $70 toilet that treats urine and excrement by mixing it with charcoal to produce highly fertile soil.8 The scandal of world famine and lack of access to clean water in the undeveloped countries of the world that can be delivered through rather straightforward engineering is exacerbated 100-fold when contrasted with the ingenuity of the advanced engineering in the developed world for the health and well-being of its citizens. A good example is Ben’s saviour—the heart pacemaker—medical engineering par excellence. The evolutionary events of this story occurred over seven uneven phases spaced over two millennia.9 The story is worth following in some detail since it illustrates how engineering ingenuity can deliver the most remarkable advances—as long as the willingness and resources are available to turn a dream into a reality. The net result is another example of the creative artfulness of a functional aesthetic.

Pacemaking The whole pacemaking story is based on electrotherapy—using electricity to stimulate human tissue. A pacemaker has a battery that creates the potential that drives a flow of current through an electric circuit. That circuit can detect changes in the electrical field created by the electrical waves moving through the heart and can deliver electrical pulses to the heart to make the heart muscles contract if required. Getting to this kind of sophisticated engineering required relentless curiosity, much frustration and lots of trial and error. Progress occurred in fits and starts but then with rapid development over the last fifty years largely because of the availability of the transistor and integrated circuit. A major medical driver was that paediatric open-heart surgery to repair congenital heart defects was too often leading to heart block—a disruption of the electrical pulses that control the beating of the heart—even in well-performed operations. The story begins in Ancient Greece. Hippocrates, author of the famous medical oath, was the first to note that ‘Those who suffer from frequent and strong faints without any manifest cause die suddenly’. Roman physicians are known to have treated patients in pain with electric rays and other electrically charged sea creatures. Much later, in 1600, Englishman William Harvey restarted a pigeon’s heart with a simple flick of the finger. Bradycardia (slow pulse rate) was probably first described by a Slovenian, Marcus Gerbazius in 1717. A Dane, Nickolev Abildgaard, put electrodes on the sides of a hen’s head in 1775—and it fell apparently dead—but then he applied the electrodes over its chest, and it recovered. In 1791, an Italian, Luigi Galvani published experiments on electrical phenomena in frog muscles and frog hearts.

172

8 Well-Being

After Galvani’s experiments, it was becoming clear that electricity had a pronounced effect on the heart. There was no real understanding of what was going on but by the nineteenth century rudimentary forms of electrical stimulation were being used to treat cardiac disease. The Irish surgeon Robert Adams linked apoplexy with disorders in cardiac rhythm in 1827 as did another Irishman William Stokes in 1846. Heart rhythm problems that led to a sudden loss of consciousness became known as the Stokes–Adams syndrome. Around 1880, a Prussian lady, Catharina Serafin went to Hugo Von Ziemssen’s clinic with a chest tumour which had been removed by an unknown person. The operation must have been rather dramatic because it is reported that part of her chest wall had been removed and her heart could be seen through a thin layer of skin. Von Ziemssen stimulated her heart with an electric current and reportedly changed her heart rate at will. John MacWilliam was an English doctor. In 1889, he applied electricity across the chest to ‘excite rhythmic contraction… to stimulate by direct means the action of the heart which has been suddenly enfeebled or arrested in diastole (the phase when the heart refills with blood) by causes of a temporary or transient character’. Karl von Vierordt was German and the first to measure blood pressure with an instrument he called a sphygmograph. It used a system of levers and weights to stop blood flow in a similar way to a modern cuff. Then Karel Frederik Wenckebach, a Dutchman, used it in 1899 to detect and describe atrioventricular block, the impairment of electrical conduction between atria and ventricles. The first version of an electrocardiogram was made in 1887 by Augustus Desiré Waller. Gabriel Lippmann had created a device in 1873, called a capillary electrometer, to detect small electric currents and Waller used it to obtain a wavy line or trace of the heartbeat. Waller learnt that ‘each beat of the heart gives an electric change, beginning at one end of the organ and ending at the other’. He was convinced that he could measure these ‘electromotive properties of the heart’ from the skin surface and proceeded to do so with the electrometer connected between the left and right hands or between the front and back paws of his pet bulldog, Jimmie. Apparently, a question was raised at the House of Commons concerning this ‘cruel procedure’ under the Cruelty to Animals Act of 1876 but Jimmie seemed to be unaffected. The clinical significance of the electrocardiogram was not recognised at the time. Waller said: ‘I do not imagine that electrocardiography is likely to find any very extensive use in the hospital’. Willem Einthoven was a key figure in our story. He used Lippmann’s device to trace the four points of the heart trace (Fig. 8.6). He called them A, B, C, D but later he changed the labels to P, Q, R, S and T, which how they are described today. In 1902, he made the first direct recording of the human electrocardiogram using a string galvanometer—an instrument for detecting and measuring small electrical currents. He hung a long thin silver-coated filament vertically between electromagnets. The patient sat with both arms and left leg in separate buckets of saline solution (Fig. 8.7). The three buckets acted as electrodes to conduct the small electric currents, generated by the patient’s heart, to the filament which then moved sideways. The shadow of the filament made a trace on a moving photographic plate.

Pacemaking

173

Fig. 8.6 Stages of the PQRST cycle of the heart

Fig. 8.7 One of the first commercial attempts at electrocardiographs 1911—the hands and foot are immersed in a salt solution. Public domain via Wikipedia Commons

Medical doctors now refer to the three electrodes as an Einthoven triangle—the basis of the modern system that uses 12 leads. The three Einthoven leads are labelled I, II and III in Fig. 8.8a. They and the other 9 leads in a modern system are each ‘looking at’ the complex electrical signal coming from the heart from a different perspective. In other words, each lead captures a different ‘view’ of the flow of electricity through the heart. Each view is a vector. A vector is something that has a magnitude in a direction. Scientists usually show a vector as an arrow—the length is the magnitude

174

8 Well-Being

Fig. 8.8 a The leads I, II, III for the Einthoven Triangle. b The Einthoven Triangle with cardiovectors for the R wave in the walls of the ventricle

of the vector and the orientation is its direction. In Fig. 8.8b, the length of an arrow is the voltage pointing in the direction of the arrow. In other words, the vector arrow represents a summary of a complex changing wave as it travels through the heart during each heartbeat. Figure 8.8b is called the electrical Einthoven triangle—it is not the triangle of leads in physical space as in Fig. 8.8a but rather an electrical model that has the heart at its centre. The full technical explanation is beyond the scope of this book. In Fig. 8.8b, we can see how the R part of the heartbeat is captured by each lead. Lead I captures a horizontal signal from negative to positive. Leads II and III capture the signal both at 60° to the horizontal and travelling towards the left leg. These individual vectors in each lead are combined as shown to make the resulting vector shown on the heart at the centre. Each of the stages of the PQRST wave is shown in Fig. 8.6. The electrical wave starts from the Sinoatrial (SA) node—the heart’s natural pacemaker—located at the top right of the right atrium. In the diagram, at each stage of the cycle PQRST, the wave is again represented by an arrow as a vector. The magnitudes of the vectors are different at each stage and they point in a different direction. The P wave—pointing down and right—causes the muscles of the atria to compress and squeeze blood into the ventricles. When the wave reaches the atrioventricular (AV) node it pauses as the ventricles fill with blood. The wave then continues its journey down the septum (the wall between the two ventricles) as the Q wave. It then travels on to the walls of the ventricles (R, S) causing them to squeeze blood out to the lungs and the rest of the body. Finally, the ventricles relax (T). The leads are connected to an electronic box which amplifies the signals and sends them to an oscilloscope or printer to display them. Einthoven was awarded the Nobel Prize for Physiology and Medicine in 1924. Einthoven used only 3 electrodes—the other 9 in a modern system were added to help gain even more detail through work by American engineers Frank Wilson in 1933 and E Goldberger in 1942. The first cardiac pacemakers were independently developed by Australian Mark Lidwell and American Albert Hyman. In 1928, Mark Lidwell saved the life of a child

Pacemaking

175

by inserting a needle and alternating current into the patient’s ventricle. Albert Hyman described his artificial pacemaker in 1932. He had previously injected a stimulant drug into the heart but realised that it was not the drug that was important but the needle. His device used a direct current (DC) generator powered by a hand-cranked motor to direct electrical impulses into the right atrium. He used a bipolar needle— an electrode with two insulated wires. He delivered impulses at rates of 30, 60 or 120 per minute. Unfortunately, his work was frustrated and eventually derailed by technical problems and the attitude of other doctors. His pacemaker was dismissed as gadgetry. The Journal of the American Medical Association did not even report his experiments. No one would initially manufacture his device although a batteryoperated version was eventually manufactured by Siemens-Halske in Germany and their American subsidiary Adlanco. It was called the Hymanotor and unfortunately found to be rather ineffective when tested in practice. The first mains-powered pacemakers were developed in the early 1950s and were bulky because they contained large electrical valves the size of light bulbs. They were wheeled around on carts and plugged into wall sockets. In 1949 Toronto, Canada, Wilfred Bigelow and John Callaghan started to experiment on stimulating the sinus node—the natural pacemaker of the heart. They were operating on a dog when its heart stopped suddenly. Bigelow reflected ‘Out of interest and in desperation, I gave the left ventricle a good poke with a probe I was holding. All four chambers of the heart responded. Further pokes clearly indicated that the heart was beating normally with good blood pressure’. They recruited John Hopps, an electrical engineer. Hopps designed what was perhaps the first electrical device specifically built as a cardiac pacemaker. The electrical impulses were transmitted through a vein to a bipolar catheter electrode inserted into the atria. Atrial pacing was readily achieved, and heart rate was controlled with no uncomfortable chest wall contractions. In 1951, in Boston, Paul Zoll read of the work done by Bigelow, Callaghan and Hopps. He developed an external table-top pacemaker and used it to treat heart block. Zoll’s pacemaker was an electrocardiograph to monitor cardiac rhythm and an electric pulse generator to pace the heart. The pulses were delivered through a pair of metal electrodes strapped to the patient’s chest directly over his heart. Unfortunately, the electrodes irritated the skin and the repeated electric shocks were painful. The equipment was also bulky and heavy and could only go as far as the mains lead would allow. What was needed was something much less bulky, portable, and ideally, wearable. The American electrical engineer Earl E. Bakken produced the first batteryoperated wearable pacemaker. Earl and his brother-in-law Palmer Hermundslie had co-founded a company called Medtronic in 1949 in a garage in northeast Minneapolis. For the first few years, they repaired hospital electrical equipment. Meantime surgeon Walton Lillehei was performing many open-heart operations to repair congenital defects. Unfortunately, around 1 in 10 patients developed complete heart block. Drugs helped in the short term, but another solution was needed. Lillehei and his co-workers made a multi-stranded stainless-steel wire in a Teflon sleeve and implanted one end directly into the muscular wall of the heart and the other end to a stimulator and they buried another electrode under the skin. They found that they

176

8 Well-Being

could pace the heart with only 1.5 V. In 1957, a 3-year old girl had heart block after an operation, so they used their new pacemaker and her heart regained rhythm, and she survived. There was no rejection and no damage to the beating heart and the wire was easily removed. The equipment was still far from ideal and further problems soon became apparent. The stimulator was large and heavy and rather overwhelming, especially for children. But there was one final and fatal flaw. On 31 October 1957, there was a three-hour power failure and a baby died. The problem was that the hospital had emergency power back up for surgical and recovery areas but not in its wards. Lillehei asked Bakken if Medtronic could come up with something better. Transistors were at that time becoming widely available. Bakken recalled seeing a circuit for a transistorised metronome in the April 1956 back issue of a magazine called ‘Popular Electronics’. The circuit transmitted clicks through a loudspeaker and the rate of the clicks could be adjusted to fit the music. Bakken modified the two-transistor circuit, put it—without the loudspeaker—into a four-inch-square and inch-and-a-half thick aluminium box with terminals and switches on the outside. He powered the circuit using a miniature 9.4 V mercury battery. He provided an on–off switch and knobs to control the stimulus rate and amplitude. His device was tested on a dog and it worked, and the next day was used successfully on another little girl. The first production run was for ten units. The product literature was rather bold: ‘So small and light that it may be attached to and worn by the patient, the Medtronic Cardiac Pacemaker stimulates ventricular function in cases of atrioventricular dissociation that are induced during the surgical repair of septal defects, or that occur spontaneously as in Stokes-Adams syndrome’. Bakken’s pacemaker was one of the first successful applications of transistor technology to medical devices and helped to launch the new field of medical electronics. The first pacemaker to be implanted was in Sweden in 1958. Ake Senning and Rune Elmqvist inserted a device into a 43-year old engineer called Arne Larsson. Larsson had to be resuscitated many times every day—his situation was so hopeless that the operation was simply a desperate attempt at rescue. Ake Senning knew of Lillehei’s work on external pacemakers. Rune Elmqvist was a medical graduate who had become an engineer. These two men began to collaborate closely in 1950 and saw that the main problem with external pacemakers was infection along the lead. They decided to design a fully implantable system (Fig. 8.9). The first attempt worked for only a few hours, but they tried a second one which lasted for about a week. The story had a happy ending for Arne Larsson he survived both of his saviours. In total, he had five lead systems and 22 pulse generators of 11 different models until his death on December 28th, 2001 aged 86 of a totally unrelated condition. In 1958 Wilson Greatbatch, an electrical engineer at the University of Buffalo, was working on the design of an oscillator. He was trying to record tachycardias— episodes of rapid heartbeat—but he accidentally discovered the way to make an implantable pacemaker. He misread the colour coding on a resistor and inserted the wrong one into his trial design. He took his discovery to surgeon William Chardack. Chardack and Andrew Gage used the device on the heart of a dog, and it worked. ‘Well, I’ll be damned!’ Dr. Chardack was heard to exclaim. Later Greatbatch wrote

Pacemaking

177

Fig. 8.9 Diagrams of artificial pacemakers and leads into the heart. Image by Npatchett. CC BY-SA-4.0 via Wikipedia Commons

‘I seriously doubt if anything I ever do will give me the elation, I felt that day’. Greatbatch patented his invention in 1959 and Chardack reported the first success on a human when in 1960 he implanted a device in a 77-year old man with complete heart block. The patient survived uneventfully for 2 years before his death from natural causes. In 1961 the team reported successful implantations in 15 patients. A Czech, Barouh Berkovits, worked for Medtronic in 1964. He designed a demand pacemaker which is the basis of most modern pacemakers. The first pacemakers with microprocessors appeared in the 1990s. They can detect and store information about cardiac events and have in-built mathematical algorithms (rules) to deliver therapy automatically. In other words, they can modify their internal pacing parameters according to the changing needs of the patient and they can adjust to the patient’s activity levels. Modern pacemakers weigh no more than 25 g. Future developments will make them even smaller (possibly small enough to be inserted by a catheter), longer lasting (through improved batteries or by harvesting power from the patient’s own heart or body) and able to perform even more functions possibly wirelessly without the need for leads and safer in that they may be used in MRI scanning machines. Innovations in programmability and telemetry will increase the diagnostic capabilities of pacemakers. There is also the possibility that gene therapy will develop ways of turning stem cells or cardiac muscle cells into pacemaker cells.

What Next? Engineering continues to help us thrive—to deliver basic health needs of feeding our hunger, quenching our thirst and disposing of waste as well as helping us live longer

178

8 Well-Being

and more comfortable and less painful lives through pacemakers, artificial hips, drug delivery systems and tissue engineering. But we have some big challenges. The entirely avoidable famine, lack of access to clean water, the killer diarrhoea and high infant mortality in too many parts of the world could be expunged from the planet. Pollution is entirely man-made and avoidable, but it can only be eliminated if the whole world rallies around that singular purpose. It is incumbent on us to pressure our world leaders to bring these changes about by finding sufficient social engineering ingenuity. Medical engineering is in its infancy—the technical challenges are difficult but the unforeseen social and ethical complexities are formidable. Health informatics will eventually improve the acquisition, storage, retrieval and use of information in health care—you may already have a ‘wristwatch’ that measures your pulse rate. New equipment will provide masses of ‘big data’ that will be used to improve treatments. New biomaterials based on polymers, ceramics and composite materials will be developed beyond hip and knee replacements. Mechanics applied to living systems, called biomechanics, will produce new understanding. Regenerative medicine— engineering human cells, tissues or organs will improve. Medical imaging, which includes X-ray, magnetic resonance imaging, ultrasound, endoscopy, thermography and positron emission tomography (PET) will be further developed. Genetic engineering will remain controversial but potentially valuable to all those with inherited diseases—the major dilemma will be to avoid the unintended consequences of both good and bad practice. Engineers and medical doctors are realising how much they have in common as both professions are people working for people. Their close collaboration will be of enormous future benefit for well-being and health. But the grand challenges confront us. Well-being and health are only part of the bigger question of human flourishing. End Notes 1. For World Bank life expectancy data see https://data.worldbank.org/indicator/ SP.DYN.LE00.IN (Last accessed February 2019). 2. Figure 8.2 Credits Figure 8.1.1: Image in public domain from the Yorck Project: 10.000 Meisterwerke der Malerei. Distributed by DIRECTMEDIA Publishing GmbH via Wikipedia Commons. Figure 8.1.2: Image in public domain from Project Gutenberg’s Young Folks’ History of Rome, by Charlotte Mary Yonge via Wikipedia Commons. Figure 8.1.3: Image by S. Anand. CC BY-SA 2.0 Wikipedia Commons. Figure 8.1.4: From Agriculture by Anya Smart. Figure 8.1.5: Image by Biggishben. CC BY-SA 2.0 Wikipedia Commons. Figure 8.1.6: Image in Public domain via Wikipedia Commons. Figure 8.1.7: Image in Public domain from Fermi Lab history project via Wikipedia Commons. Figure 8.1.8: Image transferred from en.wikipedia.org. Author Grautbakken. CC BY-SA-3.0 via Wikipedia Commons.

What Next?

179

3. See http://www.steamploughclub.org.uk/history.html (Last accessed February 2019). 4. For evidence that by the Middle Ages in the UK people were well aware that not all water was safe to drink see http://www.medievalists.net/2014/07/peopledrink-water-middle-ages/ (Last accessed February 2019). 5. https://www.revolvy.com/main/index.php?s=Alexander%20Cumming (Last accessed February 2019) refers to Gloria Clifton (2004) Oxford Dictionary of National Biography, Oxford University Press. (Subscription or UK public library membership required). 6. The Oxford English Dictionary defines crap as something of extremely poor quality https://en.oxforddictionaries.com/definition/crap (Last accessed February 2019). 7. George, R. (2008) Big Necessity: Adventures in the world of human waste, Portobello Books, London. 8. See http://loolaboo.com/ (Last accessed February 2019). 9. The account of the development of the heart pacemaker leans heavily on an online article by Oscar Aquilina, a cardiologist working in Malta—see http://www.ncbi. nlm.nih.gov/pmc/articles/PMC3232561/ (Last accessed February 2019).

Chapter 9

Flourishing

Keep on yearning, always learning, world is burning

West Gate Bridge The only phrase I could think of was ‘There but for the grace of God go I’. It was 1971 and I was reading the specialist newspaper called ‘Construction News’. The article was an account of proceedings at a Royal Commission of Inquiry into the collapse of a span on the west side of the West Gate Bridge over the River Yarra in Melbourne Australia when 35 people were killed.1 The inquiry panel was questioning Christopher Simpson—the engineer responsible for the construction of the bridge on the east side of the river and working for the designers of the bridge called Freeman Fox and Partners. I felt keenly that they might very well have been questioning me had circumstances been slightly different because Christopher and I graduated together in 1964. I read verbatim the difficult questions and answers put to Christopher. I felt keenly that I could have been him. I wondered how I would have coped. Being part of, or witnessing, a terrible disaster is clearly very distressing. In this final chapter, we ask, ‘How do we minimize this kind of distress and continue to flourish in the future?’ How do we learn from past events? Are there patterns of unwanted events in past disasters that were lying hidden, unrecognised and unforeseen as the final failure incubated? Can we identify those patterns and use the understanding they give us to avoid future failures and to ameliorate the effects of climate change? What are the ‘grand challenges’—both natural and man-made? Are they global and/or local? How important is it to slow up, halt and reverse the fragmentation of the professions into ‘silos’? What is ‘joined-up’ thinking in ‘joined-up’ organisations and nation states and how important is it?

© Springer Nature Switzerland AG 2020 D. Blockley, Creativity, Problem Solving, and Aesthetics in Engineering, https://doi.org/10.1007/978-3-030-38257-5_9

181

182

9 Flourishing

Becoming a professionally qualified engineer is not easy—it requires the kind of hard work and dedication that generates self-respect. Almost all engineers I know are modestly proud of what they have achieved. When you have been part of the designing and making of a challenging project you have great satisfaction when you see the finished product. To be working on a bridge as part of it collapses is devastating and something that stays with you for the rest of your life. Recovery depends on a strength of character that goes beyond the ordinary. But personal trauma need not stand in the way of success. Psychologists Feldman and Kravetz2 have noted this in people who have an accident, receive a devastating medical diagnosis, lose a loved one or live in a country where war and famine are almost routine. They identified five factors that seem to help people: hope, personal control, social support, forgiveness and spirituality. They write ‘truly accepting the consequences of a trauma with realistic thinking rather than delusional positive thinking can open people up to true hope – something that enables setting and achieving goals that ultimately can improve one’s life….A realistic view of the situation + a strong view of one’s ability to control one’s destiny through one’s efforts = grounded hope’. Flourishing is about achieving your potential—individually and collectively despite obstacles and setbacks. The word stems from the Latin florere ‘to bloom, blossom, and flower’. Aristotle was the first to understand its importance. He saw that flourishing occurs when we want to do what we ought to do—take pleasure in moral action. He thought that earthly happiness is a by-product of using our individual capacities to realise our potential. In that sense what really matters is not what we believe but how we behave. Karen Armstrong3 calls the religious emphasis on belief as determining action, a metaphysical mistake and an accident of history. Acting by creating imparts feelings of achievement. As we noted in Chap. 2 personal hobbies like gardening, drawing, craftwork or simply just repair jobs around the home can be very satisfying. Throughout history we have built big cathedrals, mosques, churches and temples as expressions of communal faith. Experiencing those spaces often creates feelings of high spirituality. Sports stadia like Queen Elizabeth Olympic Park in London, exhibition halls and museums like the Guggenheim Museum in Bilbao, large-scale art works such as the i360 tower in Brighton opened in 2016 (Fig. 9.1) and the colossal Cornish Man Engine4 (Fig. 9.2) are all artistic expressions that embody the conscious use of the imagination. Given the complex challenges we are facing in the twenty-first century, we can learn a lot about our collective future from the advice from Feldman and Kravetz and the experiences of those who live through personal trauma such as the West Gate Bridge collapse. In order to flourish together, we need to maintain hope, take collective control, support each other, live with a spirit of forgiveness that unites rather than divides us and develop a sense of spirituality that goes beyond conventional religion. The five principles of Part, Unintended, Preparedness, Ingenuity and Learning are key ingredients. When we truly understand ourselves as Part of the natural world then we have hope for a better future. When we admit our tendency for technical triumphalism—past arrogance in thinking that we are in total control of the natural world then we are ready to accept unforeseen and Unintended consequences and unknown unknowns. When we recognise that Preparedness for an uncertain future

West Gate Bridge

183

Fig. 9.1 British Airways i360 viewing tower, Brighton, England Conceived and designed by Marks Barfield Architects and John Roberts (Jacobs Engineering Group). By permission of Brighton i360 Limited

(a)

(b)

Fig. 9.2 a Mineworkers riding the man engine at the 234 fathom (428 m) level of Dolcoath Mine, Camborne, Cornwall in 1893. John Charles Burrow (1852–1914) by permission of the Steve Colwill Collection. b The mechanical giant puppet built to celebrate the Cornish man engine. Image by Rod Allday CC BY-SA-2.0 via Wikipedia Commons

is by developing communities that integrate rather than divide us with a sense of overall common purpose. When we accept that the golden rule of all religions is ‘Don’t do to others that you would not have them do to you’ then we have a key to spirituality. Again, this rule emphasises that the way we act leads to the way we believe rather than, as most religions now teach, belief determines how we should

184

9 Flourishing

behave. We should be using all of our considerable ingenuity to find new ways of living to learn how to be resilient when an unknown becomes a known—for example, an unforeseen consequence (an unknown) of climate change hits us (becomes known). Christopher’s story started with an offer to work in the company that Sir Ralph Freeman senior had created. Freeman had been one of the most eminent engineers of his day—for example, he designed the iconic Sydney Harbour Bridge (1932) in Australia. His son, also Ralph Freeman (and also knighted for his services to engineering) joined the business after the war and was made a partner in 1947. He had pre-war experience of bridge design and construction in Africa and worked on the 3.2 km steel girder road and rail Storstrøm Bridge in Denmark. He served with distinction in the war as chief engineer to 21 Army Group and was responsible for military bridges in France, Belgium, Holland and Germany. Freeman junior took over as senior partner at Freeman Fox in 1963, a position he held until he retired in 1979. Freeman Fox and Partners were in the early 1970s without doubt the top bridge designers in the world. They were responsible for the Forth Road Bridge, the Severn and Wye Bridges, the Erskine Bridge and both Bosphorus Bridges in Turkey. Freeman’s other most notable partners were Sir Gilbert Roberts, Oleg Kerensky and Bill Brown. Roberts and Brown were responsible for the design of the West Gate Bridge. They had previously developed new design concepts including suspension and cable-stayed bridge decks with a stable aero-foil-shaped cross section, box girders with stiffened steel decks, new methods for spinning of cable wires and a new cable configuration. As soon as he started work, Christopher began to gain some very valuable experience of bridge design and construction—including a two-year spell in Germany. He was then seconded in 1970 to the contractor Fairfield-Mabey working in Scotland on the Erskine Bridge, and responsible for the construction of the south side. One day he had an urgent phone call asking him to go straightaway to Australia to work on the West Gate Bridge in Melbourne. A new contractor, John Holland Pty, had been appointed by the Bridge Authority. Although the company had a high reputation, it had no experience of the construction of long-span steel bridges. The contractual position had become unclear with no definition of responsibilities of the respective parties. Christopher was given the responsibility for the construction of the bridge on the east side of the river. In Melbourne, life was challenging due to demands of the job as well as settling down in a new country. But people were very friendly and helpful, and Christopher built good relations with the contractor—for example, he played squash regularly with Bill Tracey who worked for John Holland. Unfortunately, Bill was one of the 40 people killed. The other engineers who lost their lives and had worked closely with Christopher were Jack Hindshaw and Peter Crossley. Jack was the Resident Engineer. His role was to represent the designer on site—on behalf of the client— and to make sure that the contractor carried out the work as required. Jack had only been in Australia for a relatively short time after replacing the previous resident engineer. Christopher recalls that Jack was small in stature but very likeable with a strong Lancastrian accent. Christopher had worked in the London Office with

West Gate Bridge

185

Peter Crossley. Peter was Deputy Resident Engineer and an exceptionally intelligent engineer with a degree from Cambridge University and a very successful career ahead of him. The building of the bridge had not been easy with political, industrial and engineering problems. The original contractor was called World Services and Construction Pty Ltd. They had many difficulties with the trades union and eventually the company was dismissed from the contract by the Bridge Authority. As well as the problems at West Gate, Freeman Fox also had to cope with the collapse of their Milford Haven Bridge in Pembroke in Wales in 1970. That bridge differed from West Gate in that it was an all steel structure and was being erected by cantilevering out from a pier. Lengths of the full cross section of the box were welded on the end of a growing cantilever until the total length reached and rested on the next pier. The collapse occurred just as they were near to the next pier. At that point, the cantilever was at its longest and hence most vulnerable stage. The forces in the bridge at the root of the cantilever were at their maximum. As a direct consequence, a steel diaphragm at the pier buckled, the bridge effectively broke its back, the end of the cantilever fell to the ground and the span ended up broken and leaning at a sharp angle on the first pier. At West Gate, all of the long approach spans were constructed in concrete. The main span over the river and the two spans each side of the river were designed as a composite structure—a steel box girder topped by a concrete deck. This was quite unusual at the time for a bridge of this size but was used because of political pressure to have as much local content as possible. The problem was that the Australian steel mills could not produce enough steel of the required quality for a full steel design. Also, there had been problems with cracking in the steel beams of a bridge built earlier in Melbourne. Long-span bridges have to be made as light as possible because the effect of its own weight is large when compared to the effects of the traffic. The result was that the erection of the bridge was challenging. The method of erection of the first of the steel spans was quite different from that used at Milford Haven. Two full span lengths of half boxes were assembled on the ground directly alongside the span. So, if you imagine the bridge in cross section, the joint between the two halves was down the centre line of the bridge. Each half section was the length of the span. After assembly, each one was jacked on towers high into the air and moved across to the final position on top of the piers at either end. The adjacent steel plates along the centre line were then to be joined together to form the full cross section. The joint was to be made by inserting a series of friction grip bolts into pre-drilled holes on both sides of the centre line. The problem was that the half sections were asymmetrical, and the abutting plates were very flexible. Consequently, the plates buckled and distorted along their length. It was difficult to get the holes in the steel plates aligned so the bolts could be inserted and tightened. The spans on the east side of the river were constructed first—to learn lessons that could be used for the west side. On the day of the collapse, the identical steel span on the east side had already been built. Indeed, the partly erected bridge was visited by a number of dignitaries. In Fig. 9.3, Christopher is shaking hands with the Governor of Victoria Major General

186

9 Flourishing

Fig. 9.3 Meeting on West Gate Bridge. Image by Ketlar Photographic—provided by Christopher Simpson

Sir Rohan Delacombe. Christopher remembers that on the fateful day Jack and Peter came to him on the east side of the river. They had received a message that there was a problem on the west side—advice was needed but there was no indication that it was critical. Jack asked Christopher to accompany him. As they prepared to take the ferry, Christopher saw something that concerned him with the jacking operation to enable the cantilevered span to land on the next pier on his side of the river and he felt he needed to attend to it. So, he asked Jack if he could sort that out first before joining them on the west bank. Jack agreed and with Peter boarded the ferry to cross to the other side of the river. Christopher went up and on to the deck of the eastern span. A little later he was briefing some of the men when they heard a horrendous crash coming from the west bank. Christopher told me that the noise, dust and smoke is something he and the others will never forget—the western span had collapsed killing 35 men on the bridge including Jack and Peter who had only just reached the top of the deck. The Royal Commission of Enquiry uncovered all sorts of incubating issues and the trigger event. They concluded ‘There can be no doubt that the particular action which precipitated the collapse…. was the removal of a number of bolts from a transverse splice in the upper flange plating near to mid-span. The bolts were removed in an attempt to straighten out a buckle which had occurred in one of the eight panels which

West Gate Bridge

187

constitute the upper flange. The buckle in turn, had been caused by the application of kentledge (heavy weight) in an attempt to overcome difficulties caused by errors in camber. To attribute the failure of the bridge to this single action of removing bolts would be entirely misleading. In our opinion, the sources of the failure lie much further back; they arise from two main causes. Primarily the designers of this major bridge, FF & P (Freeman Fox and Partners) failed altogether to give a proper and careful regard to the process of structural design. They failed also to give a proper check to the safety of the erection proposals put forward by the original contractors, WSC (World Services and Construction Pty Ltd). In consequence, the margins of safety for the bridge were inadequate during erection; they would also have been inadequate in the service condition had the bridge been completed. A secondary cause leading to the disaster was the unusual method proposed by WSC for the erection…. This erection method, if it was to be successful, required more than usual care on the part of the contractor and a consequential responsibility on the consultants to ensure that such care was indeed exercised. Neither contractor, WSC nor later JHC (John Holland & Co), appears to have appreciated this need for great care, while the consultants FF & P, failed in their duty to prevent the contractor from using procedures liable to be dangerous’. As I later read the Commissioner’s report, I began to realize that Christopher and his colleagues on site had to deal with a set of very complex technical issues—box girders were state of the art at the time. Freeman Fox were at the top of the game technically and pushing at the boundaries of what was possible. The whole episode was really about solving a hugely difficult technical problem very severely aggravated by the political and commercial pressures surrounding the project. The role of the Bridge Authority, the human and organisational factors especially in the London office of Freeman Fox, the inexperience of the contractor and the consequential difficulties on site, had not just been circumstantial to technical matters, but had been a central part of the reason for the collapse and loss of life. In the whole of my engineering education up to that moment I had never been asked to even think about human and organisational factors. I realised for the first time that engineering is not just a technical discipline. Above all, engineering is a risky activity done by people for people. As we have seen in earlier chapters when it is done well it contributes to the quality of life. But if the people involved make mistakes that are not just narrowly technical but also include poor management practice, then lives may well be lost. In retrospect, this now sounds all too obvious but at the time it began to help me make sense of all sorts of issues deriving from the narrow technical focus of my education. It also led directly to my later work with Barry Turner and his theory of ‘incubating’ accidents-waiting-to-happen’.5

Will This Never Happen Again? How often do we hear people say after some unfortunate event, ‘We must ensure that this never happens again’? The feeling that someone close has suffered in vain is

188

9 Flourishing

very painful. We naturally want to learn the lessons. But how realistic is it to believe with absolute certainty it will never happen again. History does repeat itself—we can, and we do attempt to learn the lessons and reduce the risks down to acceptable levels—but they are never zero. We know that sloppy thinking, negligence, corruption and wrongdoing have to be rooted out. But failures incubate in many various guises and investigators may get so immersed in the details of specific events that they fail to spot the ‘big picture’ patterns that Barry Turner‘s incubating accidents model suggests. Incubation, or accident waiting to happen, as used by Turner, is an emergent property of events. Likewise, so is the actual catastrophe as the incubating preconditions are triggered by a final event. People responsible for projects need to look for signs that the pressures in Turner’s balloon are growing and when necessary take action to reduce them. In other words, they have constantly to be prepared, vigilant and careful—actively looking for evidence of incubating issues and being ready to act—to spend money to save money—especially with low-chance high-consequence risks. Some of the detailed things we learn are common sense. West Gate tells us rather obviously ‘Don’t position site huts under the bridge you are building’. But what might be the cost-risk benefit on a future restricted site where positioning the huts somewhere else is an upfront immediate real cost? Other lessons are much less straightforward and more general. For example, being prepared to admit you don’t know when that is genuinely the case risks you being seen to be incompetent. Someone else may claim to know—when perhaps the fact that they don’t know only emerges later after they have gained the advantage. People do get trapped inside their specialist groups, teams or pockets of knowledge when groups are not integrated and do not communicate well. Information or resources are not shared, and the bigger picture does not emerge—indeed sharing may threaten protection of personal information and expose vulnerabilities. Problems created by organisational silos seem to be a lesson that goes to the heart of many of the failures of the social services and criminal justice system as well as Australian bridges.6 Ultimately, learning and executing these lessons is a matter of culture—learning together has to be our collective purpose. But to do that we have to bridge some quite large cultural divides as we attempt to make sense of the patterns of behaviour that both lead to success and failure.

Revisiting Patterns Patterns are at the heart of Turner’s incubating accident model as they are for the modern game-changing disciplines of IT, quantum physics, complexity theory and systems engineering. Systems engineering is becoming a major new way of thinking because, at heart, it is about reversing the effects of fragmentation—its central purpose is integration. The way this can happen is by matching patterns of the physical world with the human and social worlds of the various disciplines of engineering. The stories of previous chapters illustrate graphically how physical systems such as ploughs, bicycles, engines and heart pacemakers have successfully evolved in

Revisiting Patterns

189

complex patterns of ingenuities entirely embedded in society and culture—notwithstanding some ‘dead ends’ and failures too. Unfortunately, the prevailing ‘western’ cultural view is that the physical and the human are distinct and disjoint. The word ‘hard’ is still frequently used to describe physical systems and the word ‘soft’ for people and social systems. Hard implies objective, impersonal, value free, cold and emotionless. Soft is subjective, personal, value laden, warm and emotional. But the notion of a hard system also suggests (especially to academic engineers and scientists) well-defined, difficult, rigorous and having optimum solutions that can be found using scientific methods. The notion of a soft system denotes ill-defined, complex, lacking rigour, with no possibility of clear solutions. In engineering, as in life, these kinds of distinctions are often sincerely held, but are also deeply unhelpful. They tend to lead to stereotypes of technological triumphalism on the one side and badinage of ‘techy nerds’ high on the Asperger’s scale on the other.7 More seriously they have in the past encouraged a profound lack of interdisciplinary understanding, sometimes active hostility, and little active collaboration between engineering and other intellectual disciplines. The formidable challenges of the future will have interwoven hard and soft system issues. They will present us with unforeseen and presently unknowable physical and social consequences including surprises and opportunities. We are much more likely to succeed if we can collaborate across all kinds of boundaries whether professional, political, gender, ethnic or religious. Systems thinking is not just a new ‘management fad’, rather it is a genuine change in outlook.8 A system is a pattern of a collection of things, whether physical or human, arranged in layers of sub-systems as per the principle of Part. Systems are chosen and defined for a particular purpose. An important aspect of the principle of Preparedness is to engineer things so that the right information (what) gets to the right people (who) at the right time (when) for the right purpose (why) in the right form (where) and in the right way (how). Easy to say but very difficult to do. Some label this way of looking at things as ‘joined-up-thinking’ because an important purpose is to harmonise the detail of a system with the big picture and the physical with the human. As a systems thinker, I start with three important ideas. First, I model my understanding of the world in levels or layers of dynamic patterns as per the principle of Part. We saw the importance of layered patterns for computers and the Internet in Chap. 6. Second, I see the components within the layers as parts and wholes at the same instance. Arthur Koestler9 captured this idea with the word holon from the Greek holos meaning whole and the suffix ‘on’ meaning a part as in proton and electron. Third I see holons as being interconnected and exchanging energy with other particular holons. In other words, holons are continuously interacting processes arranged in layers. I find that a good way of thinking about layers of holons is this. In our own bodies groups of atoms are holons that combine to become various molecules (holons) which interact in cells (holons of skin, blood, neuron cells, etc.), groups of cells combine to become tissue (holons of skin, bone, etc.) which form organs (holons of kidney, heart, brain, etc.) and systems of organs (holons of skeletal, digestive, nervous systems, etc.) that together make us who we are—as holons. We are, in effect, a massive

190

9 Flourishing

collection of organised interacting atoms—interacting simpler things that lead, in successive layers, to more complex things. Unfortunately, the systems carry a bevy of uncertainties and unintended consequences from the interactions such as disease. Disease is the principle of the unintended in action at molecular levels although of course our molecules, cells and tissue have no intentionality—that is, ascribed by us as conscious human beings in ways we do not yet fully understand. The layers do not stop with us as individuals of course. Each of us would not survive without being part of even higher level layers such as our family holons, and all sorts of social grouping holons from local societies to regional and national and international organisations. We see these layers of holons in everything. For example, a bridge is also made of atoms and molecules which form materials such as concrete, rock and steel (c.f. tissue) and we shape materials to form structural components like beams (c.f. organs) which interact to form parts of bridge-like foundations and superstructure (c.f. skeleton) which together make the bridge. But the bridge is part of a transport system such as highway (c.f. local society) which is essential to the life and economy of a region. I think that the key to achieving joined-up thinking is threefold. Firstly, an organisational culture that works through active collaboration. Secondly, common consent to work with a particular model of the structure of the interactions between the holons. Thirdly, a common view of the structure of each holon based on a variant of the decision-making processes we looked at in Chap. 3. So far, we have only covered the first two factors—collaboration and interactions between holons. The third key factor, the internal structure of the holons, can be usefully based on answers to common sets of question types required for Preparedness. Recall that the headings are why, how, who, what, where and when. The idea is that answers to why questions are the potential that drive the flow of changes in answers to questions, who, what, where and when by using methods and transformations from answers to questions how. For example, asking why questions identifies the problem in terms of needs, wants, purposes and objectives for soft systems but voltage and acceleration for hard systems. Asking who, what, where, when questions identifies possible solutions and the criteria used, to select the best one, to carry flows of change. Within that grouping questions of type who identify the people and their capabilities for soft systems in which hard systems are embedded. Questions what identify the performance indicators and state variables. Questions where define place and context and when define timings for both hard and soft systems. Finally, questions how identify methods by which inputs (values of who, what, where, when) are turned to outputs. They might include method statements, recipes but also theoretical mathematical transfer functions and computer algorithms. At the risk of being arcane, it is important to recognise that thinking of the world in this layered set of structured holons is not the reality—rather it is our way of thinking about the reality. It’s helpful because it gives us a way of coping with the huge scope, size and extent of the world so that we can model, represent share and make decisions about it in a meaningful way. It integrates our thinking about hard and soft systems with one structure and follows Popper’s three worlds in that we

Revisiting Patterns

191

Fig. 9.4 Sculpture of Aristotle at the Louvre Museum. Image by Eric Gaba, CC-BY-SA-2.5 via Wikimedia Commons

recognise that reality exists (world 1) but we can only access it through our human apparatus—our senses and our brains (worlds 2 and 3). Finding patterns across whole systems requires individual and collective ingenuity. But above all it depends upon collaboration—and that needs rigorous practical wisdom (Fig. 9.4).

Practical Wisdom and Rigour Practical wisdom is an expression of the principle of Ingenuity—being inventive, resourceful and skilful through direct practical experience. The loss of Aristotle’s idea of practical wisdom as phronesis (Chap. 3) in modern times means that we simply do not recognise, value or nurture it. Not only that, practical people are made to feel inferior compared to those with more theoretical knowledge. Of course, the roots of this loss are complex as Aristotle’s ideas have been interpreted and modified by many historical developments. For more on this see.10 Practice is often criticised as being ad hoc and lacking rigour because engineers use approximations and judgment. In fact, engineers must be, and are, rigorous in a different way for two compelling reasons. Firstly, engineering products will

192

9 Flourishing

inevitably be subject to the ultimate test—that of Mother Nature. If a bridge structure is inadequate to take the imposed forces, then it will collapse. This enforces a kind of ‘natural honesty’ and trustworthiness—a requirement that cannot be twisted by propaganda or ‘spin’. Secondly, engineers have a legal duty of care to society. Under this duty engineers must justify their decisions, if called upon to do so (e.g. when something goes wrong), in a society that questions expertise. Rigour is the strict adherence to, and enforcement of, rules to an end. Mathematical logic is the ultimate form of absolute rigour with one value—truth. It is top-down reasoning, i.e. theorems (at the bottom) are deduced from axioms (including rules) which are true (at the top) by definition. The result is a self-consistent body of true statements—but only if the axioms correspond to reality. An example of where there are differences in the axioms is Euclidean plane geometry, (that we learnt at school with axioms that apply only to 2D plane surfaces) and spherical geometry (used by navigators, surveyors and astronomers with axioms that apply to a sphere). Science is rigorous in an opposite way to mathematics because it is reasoning from the bottom up. Put simply scientists’ construct theories (at the top) and use them to deduce propositions (at the bottom) that they can test with an experiment. If successful, then they have a truth or a fact (at the bottom). Unfortunately, one fact (at the bottom) does not prove that the theory (at the top) is true in all circumstances. That is because there could be lots of other propositions deducible from the theory in many different circumstances which haven’t been tested, or just haven’t been thought of, and which could turn out not to be true. We can never be sure because there are so many possibilities. Nevertheless, the more successful tests we conduct then the more we understand the context in which the theory works and the more dependable it is to make practical decisions. Such testing is part of the duty of care of all practitioners. Practical rigour is more complex than either mathematical or scientific rigour and not well understood still less appreciated. It is the meeting of a need by setting clear objectives. Of overriding importance is the professional duty of care to deliver those objectives using all relevant and available information and expertise. Although we know that Newtonian physics may not be true throughout the cosmos, it is dependably true for most earth-bound engineering calculations, and we can responsibly and rigorously design big bridges using it. Practical rigour usually involves many values—some in direct conflict—and finding ways to reach the objectives in a demonstrably dependable and justifiable way that recognises all of our rational and emotional needs and desires. The five principles are central to practical rigour. The principle of Part is required to consider the whole as well as the parts and hence to reduce the chances of missing something important. Scientific and mathematical rigour in the scientific method breaks a problem into parts, removes most of the difficult bits that we do not know how to solve and focuses on the bits we can solve. It is therefore a process of selective inattention. Practical rigour does not have that luxury—but requires a rigour that encompasses the bits of the problem that we do not always understand too well and understanding the totality of the ‘big picture’ as well as the detail. The principle of the Unintended is essential in creating appropriate theoretical and practical models of the problem. Practitioners make sensible approximations

Practical Wisdom and Rigour

193

that respect nature but provide useful insights into possible future behaviours as well as using creative foresight to imagine what could possibly happen. Practical rigour requires diligence and duty of care that leaves no stone unturned with no sloppy or slip-shod thinking. Practical solutions must meet explicit needs and deliver a system valued in a variety of ways. Those ways are not just efficiency or cost but including such criteria as aesthetics, sustainability, practicability and resilience. Solutions have to be scanned for possible unintended consequences before being implemented, and able to be monitored to identify them during use. Because a practical system has to work for a particular period of time then the principle of Preparedness is required to be ready to monitor, fine-tune, readjust and maintain performance. That requires finding and evaluating evidence that we can depend on. That evidence needs to be testable, if possible, in as many ways as seems appropriate—for example, data from measurements will allow judgements about performance such as the displacements of a retaining wall or length of cracks in an aeroplane wing. Judgements can differ between individuals with similar levels of competence—which is why calm critical discussion is so important. Good judgment is a critical part of practical rigour. Professional opinions are not arbitrary, rather they are based on practical wisdom. Engineers test their judgements by asking themselves this question. Could we justify our decisions in court? All of this requires the principle of Ingenuity—perhaps most importantly of all for creative foresight to imagine possible unintended consequences. Practice requires the creativity to imagine what might happen—we might call it ‘imagineering’—how physical things will respond and how people might behave in future situations or scenarios. So, if we build a skyscraper how might it fail? If we write new social media software how might it be used? But engineers are human and therefore unable to foresee every possibility as 9/11 proved for skyscrapers and teenage self-harm, bullying and sexual abuse for social media. But we can learn and adapt. So finally, since judgements are decisive then the principle of Learning is also key to practical rigour—learning to improve or selfrenew. Without reflecting on and learning from experience practical problem solving does indeed become ad hoc. Practical rigour implies practical intelligence, which in turn implies practical experience. In other words, experience is necessary but not sufficient for practical intelligence—a capacity to learn, reason and understand practical matters. And practical intelligence is necessary but not sufficient for practical rigour. That is because practical intelligence and rigour require reflective learning and development on that experience. Practical wisdom, based on trust, will be required to address the grand challenges of our common future.

194

9 Flourishing

Grand Challenges Grand challenges are barriers that if removed would lead to important advances. In 2000, the UN established eight sweeping and ambitious Millennium Goals11 ranging from eradicating extreme poverty and hunger to ensuring environmental sustainability. In 2015, these have morphed into 17 Sustainable Development Goals (SDG) and 169 targets including good health and well-being, reduced inequalities and peace and justice.12 Clearly, such aspirational challenges require global partnerships involving everyone—again something easy to say but incredibly difficult to deliver. One of the first engineering organisations to ponder on more specific, yet still high level, engineering grand challenges was the USA National Academy of Engineering.13 In 2008, it issued a list of 14 grand challenges. The list largely concerns particular functions like solar power, but they are worth describing in a little detail before we consider some of the essential underlying generic engineering issues. In essence, they are contained within a more detailed expansion of the 17 UN SDGs. They are listed here in the order of the topics of previous chapters but in no particular order of precedence. Challenge number one has been recognised by many western governments in principle but action is patchy. The need is to restore and improve our infrastructure—our built environment is critically important. Infrastructure includes many of the needs covered in earlier chapters. The UK National Infrastructure Plan 2014 recognises that ‘Improving…. productivity is a vital element of…. long term economic planning. High-quality infrastructure boosts productivity and competitiveness….’ U.S. policy states that ‘efforts shall address the security and resilience of critical infrastructure in an integrated, holistic manner to reflect this infrastructure’s interconnectedness and interdependency’. Funding for infrastructure projects has been hopelessly inadequate. Too often we let infrastructure wear out before replacing it and simply exacerbate the problems—the lesson to be learned is not to neglect it through short-term thinking by moneymen. The second in the NAE list of challenges is making solar energy economical. It is the subject of ongoing research. For example, present materials for solar cells have impurities that tend to hinder the flow of electricity. Newer purer materials could reduce that whilst at the same time reducing costs. Aeroplanes like Solar Impulse will see a direct benefit. In all solar applications, power fluctuates as the weather changes so has to be stored when plentiful. Possibilities include using large banks of batteries (as Solar Impulse), pumping water and recovering energy (as hydroelectricity), employing superconducting magnets or flywheels, and electrolysing water to generate hydrogen to power fuel cells. Fuel cells have the great advantage that they produce virtually no pollution, but hydrogen can embrittle metal, so the risks of fracture need to be managed carefully. The third challenge is creating energy from fusion—the reaction in which two atoms of hydrogen combine together, or fuse, to form an atom of helium and some of the mass of the hydrogen is converted into energy. It is the process that powers the sun and the stars, and achieved in the laboratory but not, at present, practically

Grand Challenges

195

viable. The case for fusion is perhaps more speculative because currently it takes more energy to initiate and control the fusion than is produced. Fusion would be environmentally friendly with no combustion products or greenhouse gases—the actual products (helium and a neutron) are not radioactive. The next challenge, number four, is to find ways to sequester carbon. This is capturing carbon dioxide produced by burning fossil fuels and storing it safely away from the atmosphere. Manufacturers of carbonated water can dissolve pressurised carbon dioxide in water. Makers of dry ice convert liquid carbon dioxide into compressed dry ice snow. Similar approaches could be developed for coal-burning electric power plants in which smokestacks become absorption towers. But once sequestered the gas how has to be stored. Suggestions include old gas and oil fields, but they aren’t on their own large enough. Carbon dioxide with water can be severely corrosive as carbonic acid. In 2017, a number of researchers reported14 on 20 options for better stewardship of land. They identified and quantified ‘natural climate solutions’ to increase carbon storage and/or avoid greenhouse gas emissions across global forests, wetlands, grasslands and agricultural lands. They argued that the actions could mitigate around 37% of the cost-effective carbon dioxide needed through to 2030 with a greater than 66% chance of holding warming to below 2 °C. The actions also offer water filtration, flood buffering, soil health, biodiversity habitat and enhanced climate resilience. Whilst more research is needed this seems to provide a robust basis for immediate global action to improve ecosystem stewardship as part of a major solution to climate change. The fifth challenge is to reverse engineer the brain as part of improving communications and AI. Reverse engineering is the study of something to learn how it works in order to produce a copy—or an improved version. You will recall that I said that computers have been programmed to play chess rather well but are a long way from being creatively human. General-purpose artificial intelligence is a considerable challenge and not helped by media hype. Artificial brains have been designed without much attention to real ones. It’s rather like the aeronautical engineering of flying without much learning from the flight of birds. Some experts believe that they can reverse-engineer the brain and open up enormous opportunities. The security of communications in cyberspace is the next challenge. You would have to be a modern hermit not to be aware of the issues of personal privacy and national security in using social media through the electronic web of information sharing called cyberspace. These issues represent some of the most complex challenges engineers have ever faced. They range from protecting confidentiality and integrity to deterring identity theft. They are demands riddled with unintended consequences ranging from the ‘dark web’ to ‘sexting’ teenagers, self-harm and suicides. Encryption is a central issue—converting the syntax of meaningful patterns of sensitive information into code that is not meaningful unless you have a decoder. Sensitive information includes passwords, banking data, credit card numbers and other personal and private data. Unencrypted data is often called plaintext. Current methods use an encryption key to convert plaintext into ciphertext. Brute force is the crudest method to attack it—systematically trying each key until one that works is found— in that case, the strength of an encryption is directly related to the length of the

196

9 Flourishing

key. End-to-end encryption is more subtle because only the sender and receiver can decrypt it—no one else (including the Internet service or applications providers) or hacker can read it without some cryptanalysis—finding out weaknesses. Legitimate users benefit from secure encryption. Illegitimate use (for example, by terrorists or pornographers) is a threat because it is tricky for law enforcement agencies to get access. One suggestion that would be difficult to implement in practice is to divide a key into pieces so no one person or agency alone could decide to use it. Another suggestion is that a third party, such as a judge, could require a key be made available to law enforcement officers. Challenge number seven is to enhance virtual reality (VR) in communications. The aim is computer simulation for all of the senses to create the impression of being somewhere else. Virtual reality can transport you to a football game, a concert, a ship on the ocean or to a space station. With the right skills you can ride a horse, fly a jet plane, perform surgery or control a nuclear reactor. Virtual reality is not just about depicting scenes but about creating an illusion of actually being there. It attempts to recreate the actual experience, combining vision, sound, touch and feelings of motion engineered to give the brain a realistic set of sensations. One potential unintended consequence could be an effect on the mental health of some people as their actual and virtual realities merge. The eighth challenge is overwhelmingly serious—to prevent nuclear warfare and terror. We cannot assume that terrorists are not attempting to make nuclear weapons. The materials suitable for making a weapon have been accumulating around the world. Many countries have nuclear reactors—for producing power or just for doing research. Each one could potentially be used to produce the raw material for a nuclear bomb. The instructions for building devices are relatively easily attainable. If you have the materials, you could possibly make a bomb. Nuclear security is possibly one of the most acute policy issues of the twenty-first century—not just politically but also technically. The challenge is not just to locate dangerous nuclear material—but to find all of it in the world. Then to keep track of it, secure it and detect any diversion or transport that could fall into the hands of terrorists. Perhaps one of the biggest issues of nuclear safety, terrorism and modern warfare is the difficulty our business, political leaders and opinion formers have in understanding the potential impacts of AI. Separating media hype, about equipment such as drones, from the reality is not straightforward—regardless of the possibilities of unintended consequences. Leaders tend, as do many of us, to think deterministically in terms of cause and effect. Typically, the argument is ‘We think this is the problem, this is what I’ll do about it, and this is what will happen’. The difficulty in managing terrorism and AI war machines is in the last stage—what will happen is unpredictable. Terrorism is not a finite game—with a start, middle and end. Some terrorists are in an infinite game that they will continue to play even when their colleagues die. Power is no longer about being capable of destroying another nation—rather it depends on what you value. Changing regimes, assassinating enemies or persecuting war crimes won’t make you more powerful. Unfortunately, we humans tend to judge our leaders on personal strength and steadiness, and we choose the one who makes us feel secure. Limited leaders make poor choices. Civilian drones, and other yet to

Grand Challenges

197

be imagined AI machines, armed with weapons will change the balance of power. Formal declarations of war will become relics of the past. Power will be about leading your foe into doing what you want without having to fight. Leaders should be thinking about systems of checks-and-balances because the future of war is not about the military but about sustainability, resilience and protecting our civil liberties, national preparedness and resolve. Challenge nine is to manage the nitrogen cycle—processes central to producing food, controlling the impact of agriculture and sustainable development. Nitrogen is the main element in our air—nearly four-fifths of the atmosphere. We have altered the cycle in which nitrogen in the air is ‘fixed’, i.e. is converted from nitrogen gas into a form that can be used by living organisms as a nutrient. The widespread use of fertilisers and high-temperature industrial combustion has vastly increased the rate at which nitrogen is removed from the air. The resulting nitrogen oxides have caused air pollution and acid rain, polluted drinking water, eutrophication of watercourses (too many nutrients, decaying algae and depleting oxygen) and it has worsened global warming. We need to improve a range of engineering processes ranging from making fertiliser to recycling food wastes. Currently less than half of the fixed nitrogen generated by farming ends up in harvested crops. Less than half of the nitrogen in the crops ends up in our food. Fixed nitrogen leaks out of the system. Engineers need to identify the leakage points and devise systems to plug them. ‘Precision agriculture’ is one possibility—using sensors and algorithms to deliver nitrogen fertiliser as well as water and pesticides targeted to only the crops that need them. Sensors, powered with solar panels, collect data about moisture, temperature and acidity, nitrogen levels, crop yields and topography, using GPS and drones. The information is then used to manage and control resources more closely. A major difficulty is getting data from sensor to farmer in a form that is usable. This kind of approach may also be important in tackling crop failures that lead to famine. The tenth challenge is perhaps the most scandalous in the sense it should already have been done—it is to provide access to clean water for everyone across the world— a disgraceful lack of well-being that is responsible for so many deaths—some say more than war. Estimates are that about 1 out of every 6 people living today do not have adequate drinking water and, in some countries, that proportion rises to a half of the population. Many more lack basic sanitation. Consequently, poor health results, by some estimates, in nearly 5,000 children worldwide dying each day from diarrhoea-related diseases. The eleventh challenge is to advance health informatics. We have seen the contribution of engineers in medical engineering such as heart pacemakers. Health informatics is the broader use of digital electronics and computers. It ranges from the personal to the global, from keeping good medical records for individuals to sharing data internationally about outbreaks of disease. Engineers and medical doctors will need to work together to maintain healthy populations in the twenty-first century. The twelfth challenge is to engineer better medicines and better methods of delivery of drugs as we saw through the pioneering work of Robert Langer. Improved ways are required to (a) assess quickly the genetic profile of a patient, (b) collect and manage massive amounts of data about an individual in a trustworthy way and (c)

198

9 Flourishing

create inexpensive and rapid diagnostic devices such as sensors able to detect minute amounts of significant changes in the blood. Challenge number thirteen is to advance personalised learning through improved communications. Some learners are highly self-motivated and learn by exploring knowledge on their own with minimal guidance. Others might prefer a more structured approach or being motivated by external rewards or may need step-by-step instruction. Some even will resist learning altogether and show little motivation or interest. Systems have already been designed that store instructional content, deliver it to students and facilitate interaction between instructors and learners. The fourteenth and last National Academy of Engineering challenge concerns the tools of scientific discovery. A theme of this book is that knowing and doing develop hand in hand. Engineers make the tools that scientists use, and the knowledge gained helps engineers make better tools and the process spirals as we have seen in earlier chapters. There is no reason to doubt that this will continue so perhaps this is one challenge that will never be met totally. The new bioengineering discipline of synthetic biology has exciting potential for the design of entirely novel biological chemicals and systems that could prove useful in applications ranging from fuels to medicines to environmental clean up and more. The fourteen challenges of 2008 concerned specific areas of application. As if to demonstrate how opinions even amongst experts differ and how times move on, in 2013 The National Engineering Academies of USA, UK and China got together to discuss challenges under the six headings of sustainability, health, education, enriching life, technology and growth, and resilience. Then, in 2014, the UK Engineering and Physical Science Research Council (EPSRC) produced a list of seven different generic challenges that cross-cut those of the NAE. Each of the EPSRC challenges applies to every NAE challenge to a degree. The first one is about risk and resilience in a connected world. It concerns working with complex systems, recognising that the future is uncertain, and that our science is incomplete. Recognising the existence of unknown unknowns and adopting the principle of the Unintended will be decisive. The second EPSRC challenge is controlling cellular behaviour—generic to all living things and central to medical engineering for well-being. Integrating novel engineering approaches with biology and medicine to design, create and control living systems throws up some difficult ethical issues. The key to this challenge is that new approaches are needed to design devices, molecules and surfaces which can guide cells to perform a specific target function, whilst allowing for other biological functions and mitigating against systemic effects. The third EPSRC challenge is called ‘engineering from atoms to applications’— investing in the principle of Part. Again, it is cross-cutting and about understanding and working across the scales or levels of our understanding in ‘hard’ physical and in ‘soft’ human and organisational systems. It needs thinking based on holons—seeing systems as both a whole and a part at one and the same time. The development of new materials, such as composites like carbon fibres and graphene, is important under this heading. Carbon comes in a number of forms, called allotropes such as diamond, graphite and fullerene, where the atoms are bonded

Grand Challenges

199

together differently. Diamond bonds are very strong crystal structures and so very hard and not easily contaminated. Graphite is not so rigid, but it is a three-dimensional crystal that cleaves into thin sheets. Researchers at the University of Manchester in 2004 have taken this to the limit and made layers as thin as one atom—graphene—the first two-dimensional material in the world. The layers are very strong but flexible, transparent and with high conductivity so that they can be used in transistors without any doping. Potential applications are just emerging. The material is multi-functional so can be used in many ways at the same time in electronic sensors, solar cells, membranes for purifying water, biomedical devices, sports equipment, aerospace and it can be combined with other materials for specific applications. The fourth EPSRC challenge is the set of exciting opportunities set out by two of the NAE grand challenges for well-being. They are number twelve—engineering personally targeted delivery of drugs and number thirteen—advancing personalised learning. They are examples of bespoke engineering—engineering tailored to the individual. In the past century, mass production has given us good quality products from furniture and clothing to food and drugs to bicycles and cars. We have enjoyed things we might simply have not been able to afford if individually produced by a craftsperson. But it has eroded our sense of individuality and identity. People look for and value highly individual creativity in art and design. Mass production can result in uniformity—sometimes drab like houses set out in box-like estates. Modern techniques such as 3-D printing create the possibility of much more individuality. The word printing is possibly misleading. Rather it is bespoke engineering by getting personal. We will be able to produce devices and products tailored to individual needs through computer-controlled manufacturing processes and robots. The idea has been used in regenerative medicine to create living tissue with stem cells and perhaps will even extend to the creation of new living organs. The fifth EPSRC challenge is simply called ‘big data’ (Chap. 6). We produce, access, process and use more data than ever before in the history of humankind. As you will recall data is syntax and not semantic information. In other words, information is data that have been interpreted to give it meaning and that meaning can change with context. Some companies have been labelled as data rich and information poor where they collect and store large volumes of data, but they don’t interpret to make it useful. For example, companies that collect feedback from customers and then do not act on the results. A personal example is a friend whose flight was cancelled at short notice by an airline with no offer of alternative flights or accommodation and then within a few days sent an email asking for feedback about the service they provided. One engineering challenge is to improve syntactical encryption to make data secure, robust, reliable and efficient whilst recognising the ways in which it is changing. The second derives from the semantics of how to interpret data to develop understanding—our ability to grasp the significance and importance of our data and build models to predict possible future scenarios—knowledge. Thirdly, using the data pragmatically or cognitively and further to be creative with it requires insight and inventiveness, making new connections between previously disparate ideas and being resourceful in developing them—the principle of Ingenuity. Choosing between them

200

9 Flourishing

Fig. 9.5 Patterns

requires sagacity, soundness of judgement, foresight, being judicious and prudent— wisdom. Figure 9.5 diagrammatically illustrates these stages and is a development of original sketches by Hugh MacLeod and David Somerville.15 Perception is portrayed as identifying objects, interpretation gives them meaning (as colour in the diagram), understanding derives from joining them up to see connections and relationships, and ingenuity is spotting new and previously unlikely connections. Finally, wisdom is finding connecting pathways between previously unconnected objects. Is being wise the same as being smart? The word ‘smart’ is already finding its way into our vocabulary. We are hearing the words smart meters, smart infrastructure, smart homes, smart health and smart cities. Smart is having a quick, shrewd intelligence—usually implying common sense. A smart gas or electricity meter sends its readings to your supplier electronically. You may well already have one to keep track of your use of gas/electricity and how much it is costing. But your meter is only transmitting and informing you of your data. It is a long way from being smart in the sense of creating understanding, knowledge and is totally devoid of anything approaching wisdom. The future smart city is often quoted as having digital technology integrated across many services such as schools, libraries, entertainment, transportation, power and water supplies, waste management and law enforcement—but to be a really smart city the digital systems will need to be embedded in ingeniously wise human and social systems that are fully integrated by identifying cross-cutting processes.5 Smart buildings will be accessed by autonomous vehicles, incorporate solar walls and panels, have advanced computer-assisted energy management, with as much natural lighting and green space as possible, but at the same time must cherish our environmental and contextual heritage. There is a big risk here in that using a word like smart for these early applications is hype that will ultimately be self-defeating. The equipment may well do what it is designed to do but it will fail to live up to the expectations implied by the name. A better, more realistic phrase and not raising unrealistic expectations might have been ‘master meter’. Engineers would do well to

Grand Challenges

201

consider their language more carefully and exercise more restraint on the excesses of their marketing colleagues. The sixth challenge that EPSRC called supra-structures is, like NAE challenge number six, the development and operation of better integrated interconnected infrastructures—both the physical and organisational structures/facilities—for a sustainable future.

Engineers at the Heart of Society The seventh and final EPSRC challenge is to put engineers at the heart of public decision making—improving the use of technical expertise in democratic decision making, realising that the technical is embedded in the social, appreciating that technical and social boundaries hinder collaboration, learning together and inspiring more women. This is not a call for technocracy—government by an elite of technical experts. Technocracy means that leaders are chosen for their expertise and unelected. Technocracy is undemocratic. Of course, governments can and do consult experts of all kinds in making policy. However, there is often a lack of mutual understanding of the issues. For example, for many years, despite being advised to the contrary, governments in the UK did not seem to realise that lack of continuity of policy in providing and maintaining infrastructure created extra and avoidable costs. Partly, as a consequence, many engineers find the kind of decision making of politicians and civil servants rather baffling. If engineers are to be at the heart of public decision making there will have to be a change in the current culture of engineering education and professional development. There will need to be a much greater empathy, connection and integration with other disciplines especially the social sciences, law, medicine but also surprisingly perhaps with science and philosophy. James Crowden shows us that an engineering education can be a preparation for a life outside of the profession. Though usually conceived as a vocational subject, engineering, by its very nature, is an essential part of the human condition and needs to be taught in that way. Teachers of engineering should be aware that as well as teaching to engineering they can teach through engineering. The American philosopher Carl Mitcham16 has laid down a similar challenge to engineers—he calls it the challenge of engineering self-knowledge. He relates it to C. P. Snow’s assertion about a tension across the yawning gap between the two cultures of science and the humanities. In his 1959 Rede Lecture, at the University of Cambridge, Snow said that the split was a major hindrance to solving the world’s problems. Mitcham’s gap is not between two forms of knowledge production but rather between two forms that are practical action and consequences. He contrasts designing and constructing the world—by which he means engineering in its broadest sense—with reflecting on what it means—by which he means the humanities. At root, Mitcham wants a much wider debate across the gap between engineering work in changing our physical world and the broad social context with which it interacts.

202

9 Flourishing

In the spirit of a wider sense of engineering, some lawyers and economists have started to see their professional disciplines as a form of engineering rather than as a humanity, social or physical science. In his lucid account of law as engineering David Howarth, a distinguished academic lawyer and politician, has written about transactional and legislative law17 : ‘…like engineers…lawyers want to make something useful that works for their clients. They are presented with problems to solve, an undesired current state of affairs, and a desired future state of affairs with obstacles and risks lying between the two’. The objects created by lawyers are not metal or concrete or plastic but relationships between people and designed in words rather than in drawings. The forces that lawyer’s harness are not natural physical laws but human and can include the coercive power of the state. Contracts, companies, conveyances, wills, trusts, regulations, statutes and constitutions are all useful objects designed and created by lawyers and in engineering terms are devices that produce change in the world outside of the device. Likewise, Harvard economics professor Alvin Roth pointed out in 2002 that whilst the economic environment evolves it is also designed. Entrepreneurs, managers, legislators, regulators, lawyers and judges all get involved in the design of economic institutions. Gregory Mankiw, also an economist at Harvard, agrees that economists should think of themselves as engineers rather than scientists. In April 2019, a major conference on economics as engineering was held in the USA.18

Meeting the Challenges Discerning the relationships between law, economics and engineering widens our perspective of what engineering is and helps to bridge the gap between the social and the technical. This is important because that bridging is cardinal to meeting the SDGs and all of the grand challenges. It will facilitate the kind of integrated worldview needed to promote collaboration, value practical wisdom and nurture the ingenuity of creative resilience. For example, rethinking this way could be one key to finding the social ingenuity we need to address the scandalous corruption and lack of common humanity that prevents us collectively from providing clean water and prevent famine around the world. The reasons why this kind of scandal continues is not technical—it is entirely human. Rethinking could also help us to cope with the risks of the unexpected—to take advantage when the unexpected is good and to recover well when it is bad. Risk is not well understood by almost everyone—it isn’t an easy idea. One of the largely unrecognised reasons is that it is as generic and as enigmatic as truth. Both are, at one level, straightforward. A true statement is one that corresponds with the facts. Risk is a bit like betting on the chance of some future event like having an accident as you drive your car. However, as we have seen in earlier chapters, at the deepest level truth is difficult to pin-down, and context dependent—as is risk. Risk is to action as truth is to knowledge. In other words, just as truth is an attribute of what we think we know so is risk an attribute of how we think about how

Meeting the Challenges

203

we should act. Risk, perceived and interpreted in one context, doesn’t easily transfer to other contexts. For example, many people accept higher risks when driving a car than when they are flying in an aeroplane because they feel in control driving a car but at the mercy of others in an aeroplane. Most experts agree on variations of a basic definition of risk as ‘the chance that a particular set of conditions will happen in a stated context in the future’ but interpretations differ. For example, statisticians emphasise probability as chance or degree of belief, psychologists see it in terms of how people think, feel and act and engineers see it in terms of safety and performance factors. When the risky set of conditions spell danger or harm then they are a hazard. The flip side when the conditions promise a benefit then they are an opportunity. A simple hazard is a trailing wire that someone may trip over. A complex hazard is the difficult to spot incubating conditions of an ‘accident waiting to happen’ as first articulated by Barry Turner. A chance, such as 1 in 1000, is the ratio of the number of times an event may occur to the number of possible future scenarios. So, for every 1000 possibilities 1 might occur. This kind of thinking is straightforward for a simple problem like the tossing of a dice or the spinning of a roulette wheel. Unfortunately, problems that involve the way people behave are often complex because people have the propensity to do the unexpected. In some instances, anything could happen. We have to admit the possibility of unknown unknowns and utter surprises like 9/11 and the 2008 banking collapse or the 2020 coronavirus in China. We now know that even simple physical systems can, under certain circumstances, be complex. Results from new theories of ‘chaos’ show that even quite simple deterministic processes, like the rotation of a hinged pendulum, can be unpredictable. In other words, in some situations, the number of possible futures is infinite. Therefore, a chance of one in infinity is zero and meaningless. Of course we can and do produce theories that project the past into the future, including risk analyses and predictions, but we should always remember that the theories depend on context and the principle of the unintended rules risk for any other than quite simple contexts (like a roulette wheel). Paul Grundy, an Australian civil engineer, wrote of the six steps in reducing engineer risks. First, he says, know the hazards and risks. Then identify the weaknesses. Retrofit—replace, modify and upgrade systems—to plug the weaknesses and create resilience against all hazards. Plan emergency response procedures and educate the community to understand and implement the procedures. Finally, rehearse emergency responses regularly. Grundy’s pragmatic suggestions are sensible if we target our thinking on what we think are the risks. But his scheme is also helpful in that it moves us away from a focus on trying to predict risk (predictions that are necessarily flawed and context dependent) to a focus on how we manage risk whilst avoiding telling people what to do and including and collaborating with them to build consensus about what needs doing. In other words, Grundy is helping us to recalibrate our thinking away from risk analysis towards engineering resilience. Resilience is about recovery when things go wrong. One of the ways our natural world is resilient is through diversity. Systems that are not diverse, but rely on one thing, become vulnerable. When we optimise a system, we consider only a limited set of criteria—and we expose ourselves to being vulnerable to a surprise coming

204

9 Flourishing

from something outside of those criteria. Vulnerability is the opposite to resilience— a susceptibility to damage of the ‘weakest link’. The weakest link in a chain defines the total strength of the chain even if every other link is very strong. A really tough, strong boxer with a weak chin is only as strong as his chin. A vulnerable system is not robust, i.e. not strong, healthy, hardy and able ‘to take a knock’ or persist when subject to changes or perturbations and uncertain conditions. Lack of robustness through a vulnerability is especially important when dealing with high impact, low-chance surprises such as 9/11 and Grenfell Towers. Physicists define ‘resilience’ as the ability of an elastic material to absorb energy. Ecologists define it as the ability of an ecosystem to return to an original state after being disturbed, while medical doctors refer to an ability to recover readily from illness, stress, depression or adversity. A general definition of resilience is an ability to withstand or recover quickly from difficult conditions or to adjust easily to misfortune or change. In the UK government’s critical infrastructure resilience programme (Cabinet Office, 2010), resilience is defined as ‘the ability of a system or organisation to withstand and recover from adversity’. Sustainability is a capacity to endure and implies resilience. That means a system is not sustainable when it is not resilient. But if it is resilient, it may or may not be sustainable because there are other factors, such as environmental management and consumption of resources that are needed for sustainability. In other words, resilience is necessary but not sufficient for sustainability, but sustainability is sufficient for resilience. Creating a sustainable world is a social challenge that requires developments across the whole of society—economic, societal and environmental. Sustainable living can take many forms including changing our individual lifestyles to conserving natural resources to the way we live in ecovillages and smart cities, to a greater use of permaculture, green buildings and renewable energy. What is for sure in an uncertain world is the central role of engineering—but in a less fragmented form.

Fragmentation The engineering experience has been that specialisation leads inexorably to fragmentation with consequences that may be unforeseen, far reaching and enigmatic. They include a loss of cohesion across disciplines and organisations, professionals hunkering down into their professional ‘comfort zones’, bunkers or silos in which they hold their knowledge and power, and a focus on detail at the expense of overview and the ‘big picture’. They include difficulties in tackling issues that cross disciplinary boundaries and a lack of joined-up decision making across and between disciplines and silos. The likelihood of mistakes through poor communications, lack of shared information and issues falling ‘between the cracks’ can become serious. Furthermore, the inadequacies, often only realised in hindsight after an accident or failure investigation, can erode trust in expertise.

Fragmentation

205

Fragmentation occurs in many guises. For example, economic globalisation, perhaps rather unexpectedly, is causing fragmentation between nation states and exposing the impotency of individual states within nation state blocs. The result is an understandable retrenchment by those who want to ‘take back control’ and a self-reinforcing loop where the politics of nationalism lead to more and more fragmentation and more extreme forms of nationalism. One solution is to think differently about nation states by conceiving them as holons. By this thinking a nation state, considered as a whole, can create its own version of its politics as needed for internal cohesion. At the same time, a nation state, considered as a part, will recognise its interdependence with other states. It will understand the need for integrating policies at a number of higher levels of decision making to address complex issues that are commonly held. The key to success will be finding the political will to identify what is best decided at what level in the entire hierarchy of issues. The principle of subsidiarity of the EU Treaty of Lisbon 2007 is a good starting point but not a good omen. The reason is that the principle has been lauded in the abstract but almost completely ignored in practice. In brief, the principle states that those closest to the issues know them best and are best placed to address them. In other words, we should never entrust to a bigger unit anything that is best done by a smaller one. Stated more formally subsidiarity says that models of systems should be created at the lowest practical level that is consistent with delivering their purpose. Unfortunately, because political subsidiarity in the EU has not been implemented sufficiently well, many people, especially in the UK, have felt that the EU is heading in a direction that they never signed up to. They resent the interference in our national life by what they see as unnecessary rules and regulation and they wonder what is the point? Of course, all political systems contain a diversity of views. But diversity for its own sake is not the point. Rather diversity creates different points of view that are key to the critical discussion that creates resilience and enables future success. But accommodating and managing diverse views bring challenges through the many conflicting complex issues at every level of decision making.

Learning Together Requires Leadership Grundy’s schema is a typical engineer’s response, sensible, logical and efficient— it requires the principle of Preparedness, implies Ingenuity but makes no mention of Learning. Though perhaps not intentional Grundy’s scheme tends to give the mistaken impression that people who are not technically qualified have to be told what to do without consultation and discussion—reminiscent of ‘techy triumphalism’. Doing it differently just requires a slight change of emphasis. Work together in a spirit of goodwill. Argue honestly, see other people’s point of view and above all admit when you genuinely don’t know or don’t understand. I call it ‘honest disagreement amongst friends’. It is a recognition that we need to work together and not to fragment. It is a call for businesses to reconnect with society as Lord Browne, engineer and

206

9 Flourishing

one-time Chairman of BP, has written in his book Connect.19 He says that respect, authenticity and openness need to be embedded into the heart of business models. But change requires leadership. He writes ‘People tend to treat companies as they do other human beings. If we trust someone, we are likely to forgive them when they make a mistake. But if that person has a bad history, we might not even give them a chance to explain’. He continues ‘Companies have to make friends before they need them….when firms increase their interactions with societal actors from a rare basis to often, they more than triple their chance of mutually beneficial outcomes. Top companies identify important stakeholders and work with them…using techniques to determine who to see and when…avoid outsiders being confused by multiple visits….with conflicting messages and…emissaries being played off against each other….Collaborating with others makes more sense when being the first mover might put the company at a competitive disadvantage…create gains that no single player can achieve individually….collaboration on standards and regulation provides the basis for industry wide reform’. Whilst systems thinking is necessary to meet those challenges it is by no mean sufficient. Politics also needs leaders who inspire consensus, encourage collaboration and recognise interdependence to find and deliver integrated levels of purpose. Politics does not need leaders who follow blind doctrinaire ‘policy red lines’. Greg Young has written20 that leaders shape the culture of an organisation—a large factor controlling performance. He writes about companies, but his ideas also relate to the politics of nation states. He calls for a new kind of ‘transpersonal leader’ for the twenty-first century where companies (and countries) need to be nimble and agile in responding to rapid change. These leaders will embed authentic, ethical and emotionally intelligent behaviours into the DNA of the organisation (country). They will build strong, empathetic, collaborative trusting relationships and develop a sustainable performance enhancing culture. That culture has to embody a respect for various views, enable work at developing common purposes identified at various levels and find resilient ways to dissolve the boundaries between those views. A resilient political and business culture embodies an integrity that has no place for corruption, collusion and fraud and values transparency and due diligence. By this view, nation states and companies need to move away from the sometimes-brutal competitiveness of the twentieth century to the kind of collaborative behaviour that Lord Browne identifies. As he says building trust is key.

Creativity, Problem Solving and Aesthetics—Turning Dreams into Reality I set out to do three things in this book. First to show how engineers have made and still strive to make the quality of our lives better. Second to identify and explore some of the unintended consequences of the past and the ‘grand’ challenges ahead. Third

Creativity, Problem Solving and Aesthetics—Turning Dreams ...

207

to suggest some ‘grounding’ principles that may help us to guide or steer our way through a risky future. In Chap. 1, I quoted Sir Neil Cossons challenge to engineers to ‘spread the word to the rest of the world, get out of your professional bunkers and meet the people, ignite in them some of the magic of what it is that you do’. Where does the ‘magic’ of engineering lie? Cossons was not referring to magic as sleight of hand or deception or conjuring. Rather he was pointing to the allure and fascination through which engineering can be mysteriously enchanting. By that I mean pleasing, delightful and aesthetic— engineering as creating within us an emotional relationship with things—engineering as using our collective artfulness, ingenuity and inventiveness. If in doubt, just think of your mobile/cell phone and how only a generation ago that kind of performance and functionality was science fiction. Think of Ben’s pacemaker and how astonishing it is that a small gadget implanted in his chest can save his life. Think of a big bridge like the magnificent Millau Viaduct in France which certainly has a ‘wow’ factor. Think how people can talk, face to face, across oceans via satellite communications. Think of giant space rockets that have taken us to the moon and promise more space travel in the future. The magic of all these things is in their making and the way we have collectively learned to elegantly coax the forces of nature to turn our dreams into reality. This is why I have defined engineering very broadly as producing ‘things’ that we ‘need’ like quicker ways to travel and hip replacement joints and some things we don’t need, but may like to have, like a music streaming via the Internet or a holiday in a far off country. But modern engineers specialise in one of the branches of engineering—and the result is often a lack of ‘joined-up’ thinking to address ‘big picture’ issues. As scientific knowledge grew after the Renaissance, so the professions fragmented into their ‘silos’. You will recall that the six main ‘horizontal’ branches are civil, mechanical, electrical, aerospace, computing and medical. The ‘vertical’ divisions include engineering workers, technicians, incorporated and chartered engineers. However, different countries use different terminologies—in North America, the roughly equivalent term for UK chartered engineer is professional engineer. We have traced the stories behind some of the important engineers of the past such as Michael Faraday, Joseph Bazalgette, William Armstrong, Thomas Edison and Frank Whittle as well as some of the less well-known engineers of the present—many of them women. These people are just as important to our national and international culture as politicians and celebrities because engineering is at the heart of society. We rely on it every day and it contributes to our well-being—from the roof over our heads, the food we eat, the water we drink, disposal of waste (including flushing toilets), the roads and railways we travel on through to media, communications, tv and radio, and medical equipment such as scanners. We have also seen that engineering is often taken for granted and hence undervalued—for example, lists of creative occupations rarely mention science or engineering and economic progress is rarely understood as depending on new technology. Design is often presumed to be individual emotional expression through form, appearance and symbolism whilst function is prosaic and

208

9 Flourishing

delivered by the precision of science. Consequently, engineering as applying those laws is dismissed as being relatively straightforward. We have also seen that engineering has arisen out of, and is closely related to, other kinds of making, such as craft, fine art and invention. But largely unappreciated is the idea that aesthetics is more than just beauty but concerns our emotional relationship with things. To engineer is to solve problems as only we humans can but, historically, we have also underrated practical experience and overvalued academic ability to an extent that mars our educational systems. The Roman Vitruvius understood this when he articulated a set of objectives for practice as firmness (integrity or soundness of form), function (looking for simplicity in complexity) and delight (grace as elegance of form). Buckminster Fuller expressed the aesthetic of a structure as being ‘where the stresses and strains are at ease’. What he was saying is that the aesthetic of function is an elegance of flow of energy. For example, in the way the pressure, volume, velocity and temperature of gas are manipulated in a gas turbine engine to create the thrust that propels the aircraft. The way the complicated behaviour of the Internet emerges from the layers of interacting simplicity. Modern engineered things did not suddenly appear ‘fully formed’. They evolved over time as we have seen through the stories of the plough, bicycle, engines of various kinds, equipment that relies on electricity such as radios, computers, mobile/cell phones and heart pacemakers. The common myths about engineering include the idea that engineers simply apply science. This derives from the caricature of science as the ‘laying bare’ of the inviolable nature rather than the creative result of individual inspiration to model what we experience. Science does not have all the answers we need—it is incomplete. We can expect our buildings to be ‘as safe as houses’ as long as we understand that does not imply perfection. There is always a risk that any engineered thing can fail. When things do fail because of negligence or evident lack of duty of care then those responsible should be brought to justice. But there will be many events where no one is to blame. We have also seen that all is not rosy. We have made nuclear weapons, missiles, remotely controlled drones and created the potential for cyberattacks on our information systems by hostile powers—including terrorists. We have polluted our environment and altered our climate. The global ‘grand challenges’ are many but include risk and resilience, bioengineering, energy policy, new materials and bespoke engineering. If we are to meet these challenges, then it is imperative, therefore, that we learn from the ways engineered things fail. Engineers need to break down their professional silos by adopting ‘joined-up’ systems thinking. They need to more readily see the ‘big picture’, the context, as they contemplate the detail. Whilst we can never ensure that failures ‘never happen again’ we can reduce the risks to acceptable levels, and we can design our systems to be more resilient. To do that we need to spot the ‘patterns’ as they incubate. I have suggested five grounding principles that have been, in my view, implicit in the best engineering of the past and missing in the worst, and from which we would benefit if they should become explicit in the future. You will recollect that they are PUPIL, ‘we are Part of a world of Unintended consequences for which we need to be Prepared through Ingenuity and Learning. PUPIL tells us

Creativity, Problem Solving and Aesthetics—Turning Dreams ...

209

that we are of this world not just in the world and so we are both parts and wholes. It points out that everything we decide and act on has unintended consequences—some good and some bad. It helps us to recognise and hence deal with unknown unknowns through contingency planning. PUPIL cautions us to be prepared by developing resilient and agile systems that take us away from our current focus on prediction to thinking through how we manage outcomes, no matter how unlikely, that avoid misplaced ‘triumphalism’ of human technological progress. It advises us to revalue and nurture the lost art of practical wisdom to reduce the ‘ingenuity-gap’ between those who adapt well to complex changes and those that don’t and the chasm between academic and vocational education. PUPIL admonishes us if we are not pupils in life, i.e. we don’t ‘learn from our mistakes’—something we constantly avow to do but often fail to implement. It warns us that engineering education has placed little or no emphasis on the history of its disciplines—with the consequence that many of today’s engineers see little value in their heritage.

Engineering Is a People Profession Most importantly of all PUPIL emphasises that engineering is not about things but is about people—making life better for us all. As I said in the Preface, the USA National Academy of Engineering has a campaign called ‘Changing the Conversation’. It has four messages: engineers make a world of difference, engineers are creative problem solvers, engineers help shape the future and engineering is essential to our health, happiness, and safety. Engineering is done by people for people to improve the human condition. Engineers are agents of change—they create technology by configuring patterns of flowing energy to achieve a desired outcome, a human purpose. There is a strong case for engineering to be conceived as a much wider discipline than it is currently the case. Politics, law and economics are forms of engineering—helping to engineer the flourishing of life on earth. We need engineers as never before because unexpected and unforeseen surprises will become the new norm. We must prepare ourselves to face future challenges that we cannot yet anticipate. The UN Secretary-General António Guterres has pointed out that the Barry Turner incubation of the issues around climate change is well underway (although I have no way of knowing if he is aware of Turners work) when he said ‘The world reached several dire milestones in 2017’.21 He calls for us to ‘let some of the air out of the Turner balloon’ if we are to avoid some of the extreme events that threaten our futures when he says ‘….we continue to see huge investments in unsustainable infrastructure that lock in bad practices for decades. As many have pointed out, the Stone Age did not end because the world ran out of stones. It ended because there were better alternatives. And the same applies today to fossil fuels. Our problem is not that we do not know what to do – it is how quickly we can do it’.21 A new breed of engineers is required who can read and interpret the signs and change the conversation by initiating, discussing and debating with society at large the technical challenges and opportunities that lie ahead. A new

210

9 Flourishing

perspective on engineering by non-technically trained people could give greater value to the contribution that engineers make. The media should either appoint engineering correspondents or change their technology and their science correspondents to report on engineering as a whole (rather than their current tendency to focus on digital systems). Any form of triumphalism and hype is unhealthy—but it can only be countered by the better understanding of non-specialists. At root the message is that despite all the advances in science, we know less actually than we think we know. Contingency planning that expects surprises must be the new norm. Learn anew how to learn together is the new wisdom. Use the golden rule ‘do not do to others that you would not have them do to you’, whatever you are told to believe is the tenet we should always hold foremost in our thoughts. Behaviour is more important than belief—imperfect doing is better than uncertain knowing. But to do all of this collectively requires leadership that we can trust—perhaps the biggest challenge of all? All this because engineering is really about people and their aspirations and not simply things and their functions. End Notes 1. 2. 3. 4.

5. 6.

Report of the Royal Commission into the failure of the West Gate Bridge, (1971), Victoria, Australia. Feldman, DB., Kravetz, LD (2014) Supersurvivors: The surprising link between suffering and success, Harper Wave, London. Armstrong K (2009) A metaphysical mistake, https://www.theguardian.com/ commentisfree/belief/2009/jul/12/religion-christianity-belief-science. A man engine was a mechanism of reciprocating ladders and platforms for miners to travel up and down between working levels during the nineteenth century. It was invented in Germany but used in the tin and copper mines in Cornwall until the beginning of the twentieth century. The Cornish Man Engine puppet is a 10-metre-high giant mechanism built in 2016 to mark the 10th anniversary of the Cornwall and West Devon Mining Landscape becoming a World Heritage site. It toured from Tavistock to Geevor Tin Mine in July and August 2016. Blockley D I (2020) Building Bridges between theory & practice, World Scientific Publishing, London. The events at West Gate Bridge and failures in social services and criminal justice systems seem, at first sight, to have little in common. Closer examination reveals that lack of communication across organisational silos is common to all. The murder of Victoria Climbié shocked the nation in 2000. Harold Shipman was a family doctor who was found guilty in 2000 of murdering many of his patients. In 2002, Ian Huntley was found guilty of the Soham murders. A policy briefing by the UK Local Government Information Unit in 2003 said ‘Despite all the initiatives for joint working and local partnerships, a silo mentality still appears to persist in government departments’. In all of these cases, a lack in the joining-up of agencies led to disastrous consequences. Pieces of evidence,

Engineering Is a People Profession

211

considered in isolation, were pieces of a jigsaw. Had the pieces been put together, then a very different picture would have emerged. 7. Acker, F. (2005) Autism and Engineers—is there a connection? Ingenia, Issue 25, December 2005. See http://www.ingenia.org.uk/Ingenia/Articles/343 (Last accessed February 2019). 8. Blockley, DI., Godfrey, PS. (2017), Doing it Differently (2nd Ed), ICE Publications, London. See also http://myengineeringsystems.co.uk/ (Last accessed February 2019). 9. Koestler, A. (1967). The ghost in the machine. Picador, London. 10. See https://blog.oup.com/2014/07/practical-wisdom-vsi/ (last accessed February 2019) However, we should remember that Aristotle lived in very different times and we cannot simply apply his ideas directly—but there are observations that are worth considering in their own right as well as helping us to understand how present-day attitudes have arisen. 11. In 2000, world leaders adopted the UN Millennium Declaration and committed their nations to a new global partnership. They set 8 targets as follows: 1. Eradicate Extreme Hunger and Poverty. 2. Achieve Universal Primary Education. 3: Promote Gender Equality and Empower Women. 4: Reduce Child Mortality. 5: Improve Maternal Health. 6: Combat HIV/AIDS, Malaria and other diseases. 7: Ensure Environmental Sustainability. 8: Develop a Global Partnership for Development. 12. The Sustainable Development Goals (SDGs) were agreed in 2015 following on from the UN Millennium Goals with a deadline of 2030. They are 1. End poverty in all its forms everywhere. 2. End hunger, achieve food security and improved nutrition and promote sustainable agriculture. 3. Ensure healthy lives and promote well-being for all at all ages. 4. Ensure inclusive and equitable quality education and promote lifelong learning opportunities for all. 5. Achieve gender equality and empower all women and girls. 6. Ensure availability and sustainable management of water and sanitation for all. 7. Ensure access to affordable, reliable, sustainable and modern energy for all. 8. Promote sustained, inclusive and sustainable economic growth, full and productive employment and decent work for all. 9. Build resilient infrastructure, promote inclusive and sustainable industrialisation and foster innovation. 10. Reduce inequality within and among countries. 11. Make cities and human settlements inclusive, safe, resilient and sustainable. 12. Ensure sustainable consumption and production patterns. 13. Take urgent action to combat climate change and its impacts. 14. Conserve and sustainably use the oceans, seas and marine resources for sustainable development. 15. Protect, restore and promote sustainable use of terrestrial ecosystems, sustainably manage forests, combat desertification, and halt and reverse land degradation and halt biodiversity loss. 16. Promote peaceful and inclusive societies for sustainable development, provide access to justice for all and build effective, accountable and inclusive institutions at all levels. 17. Strengthen the means of implementation and revitalise the global partnership for sustainable development.

212

9 Flourishing

13. National Academy of Engineering. Grand challenges for engineering. See http://www.engineeringchallenges.org/ (Last accessed February 2019). 14. Bronson W. Griscom, Justin Adams, Peter W. Ellis, Richard A. Houghton, Guy Lomax, Daniela A. Miteva, William H. Schlesinger, David Shoch, Juha V. Siikamäki, Pete Smith, Peter Woodbury, Chris Zganjar, Allen Blackman, João Campari, Richard T. Conant, Christopher Delgado, Patricia Elias, Trisha Gopalakrishna, Marisa R. Hamsik, Mario Herrero, Joseph Kiesecker, Emily Landis, Lars Laestadius, Sara M. Leavitt, Susan Minnemeyer, Stephen Polasky, Peter Potapov, Francis E. Putz, Jonathan Sanderman, Marcel Silvius, Eva Wollenberg, Joseph Fargione (2017) Natural climate solutions, Proceedings of the National Academy of Sciences Oct 2017, 114 (44) 11645–11650. See https:// www.pnas.org/content/pnas/114/44/11645.full.pdf (last accessed April 2019). 15. Figure 9.5 was inspired by an illustration by David Somerville based on an original by Hugh McLeod—see https://random-blather.com/2014/04/28/ information-isnt-power/ (Last accessed February 2019). 16. Mitcham, C. (2014) The true grand challenge for engineering: self-knowledge, Issues in Science and Technology, Volume XXXI Issue 1 See http://issues.org/ 31-1/perspectives-the-true-grand-challenge-for-engineering-self-knowledge (Last accessed February 2019). 17. Howarth, D. (2013), Law as Engineering, Edward Elgar, Cheltenham, UK. 18. Duarte, P., Giraud, Y. (2019). HOPE 2019: Economics and Engineering: Institutions, Practices, and Cultures, April, The Center for the History of Political Economy, Duke University, USA. 19. Browne, J., Nuttall, R., Stadien T. (2015). Connect: How companies succeed by engaging radically with society. W H Allen, London. 20. Young G. (2016), Women, naturally better leaders for the 21st century, Transpersonal leadership series: White Paper 2, LeaderShape, Routledge, Taylor & Francis Group. See https://www.crcpress.com/rsc/downloads/WP-TL2-2016_ Transpersonal_Leadership_WP2_FINAL.pdf (Last accessed February 2019). 21. The UN Secretary-General António Guterres continued ‘The world reached several dire milestones in 2017. The economic costs of climate-related disasters hit a record: $320 billion. Energy-related carbon dioxide emissions rose 1.4 per cent, to 32.5 gigatonnes—a historic high…. And I am beginning to wonder how many more alarm bells must go off before the world rises to the challenge. See https://www.un.org/sg/en/content/sg/press-encounter/2018-03-29/ secretary-generals-press-encounter-climate-change-qa (Last accessed February 2019).

Glossary

AC Alternating current in which the direction of the current varies usually as a sine curve. AM Amplitude modulation where radio signals are transmitted by carrier waves whose amplitude (size) is modulated (change or modified) by an input signal c.f. FM. Architect Architects design the form of a building to fulfil the needs of people and how it will look aesthetically. Architectural design is about the sense of space, occupancy by people, symbolism and relationship to its setting. Art Something of special or more than ordinary significance Art comes with many different forms and structures of matter such as paint on canvas, carving of wood or stone or even a large structure. As-is distinction The difference between a contextual theory or model being used as if it is true and whether a theory actually is true. Atrial fibrillation Irregular contractions of the heart muscle. Bandwidth The smallest range of frequencies within which a particular signal can be transmitted without distortion, or the speed or capacity of data transfer of an electronic system. Binary Consisting of two. Binary counting is based on 2 units (0, 1) whereas decimal counting is based on 10 units (0 to 9). Calories Units of energy. Small calories (cals) are about 4.2 joules. Large calories (Cals) are 1,000 cals. Capacitance The property of being able to store potential. Civil engineer First used in contrast to military engineer. Now it applies to construction and infrastructure work. Complexity Complex systems are difficult to describe and predict because they have many interconnected parts. They are not just complicated but may be incomplete with emergent properties from interdependencies that are unknown and unforeseen. They often cannot be ‘solved’ rather they have to be managed to desirable outcomes through collaboration. Craft Art, trade, occupation or hobby that requires special skills. Damping A mechanism through which energy is dissipated. © Springer Nature Switzerland AG 2020 D. Blockley, Creativity, Problem Solving, and Aesthetics in Engineering, https://doi.org/10.1007/978-3-030-38257-5

213

214

Glossary

DC Direct current in which the direction and magnitude vary only slightly. Design To form or conceive. Determinate structure A structure in which we can find all of the internal forces using equilibrium. Duty of care A legal obligation not to be negligent. Electrocardiogram A record of the changes in electrical potential in the heart. Emergence A situation where an attribute of an object as a whole emerges from the interactions of its parts. It is the reason why a whole is more than the sum of its parts. For example, you have an emergent property of being able to talk and walk whereas none of your individual parts can do that on their own. Energy A capacity to do work—potential or kinetic. Engine cycle Engines have to provide continuous power—so they work through a series of thermodynamic processes that cycle round and round—a thermodynamic cycle that repeats itself. Engineering Turning dreams into reality—done by people for people to improve the human condition. Engineers use science to create technology. Evolution The process by which different kinds of systems are believed to have developed from earlier forms. Evolution by natural selection is the cornerstone of Darwinian evolution. The evolution of engineering products is through intelligent development from earlier forms where the selections are made by us—‘human selection’. Field theory A theory about a region of space in which any object at any point is influenced by a force—gravitational, electrical or magnetic. Finite element method A method in which a system such as a structure is divided (in a theoretical model—not in reality) into discrete finite elements or pieces in order to find internal attributes such as forces. The elements may be any size but are usually either a whole (for example, a structural member such as a beam) or imaginary divisions (for example, of a plate, wall or floor) into simple shapes such as triangles and rectangles. Flexibility The amount by which a bar or material may stretch or squash under a force. Flow Movement as in a stream. Fluid A liquid or gas that flows and changes shape when acted upon by a force. Flutter A form of inductance and a kind of self-reinforcing dynamic oscillation. FM Frequency modulated where radio signals are transmitted by carrier waves whose frequency (rate) is modulated (change or modified) by an input signal c.f. AM. Force An object that changes the velocity of a physical body. Form The external shape of an object. Frequency The number of occurrences of a repeating event per unit time, for example, the number of oscillations or cycles occurring in a unit of time. Function The purpose for which an object is designed to perform or the role of a person. Grand challenge A big issue that requires a significant effort to resolve.

Glossary

215

Heat Energy transmitted from an object at a higher temperature to one at a lower temperature. Holistic The idea that the whole is more than the sum of its parts. Holon An object considered, at the same time, to be both a part and a whole. It has emergent properties that derive from the cooperation of the parts. Imagineering A term coined by Alcoa Aluminium in the 1940s to capture the idea of letting your imagination soar and then engineering it down to earth to make it practical. It later became the registered trademark of the Walt Disney Company and the research and development arm responsible for the creation design and construction of their theme parks. Impedance The opposition to a flow—capacitance, inductance and resistance. Incompleteness The idea that something, known or unknown, is lacking. Indeterminate structure A structure in which we cannot find all of the internal forces using equilibrium—we have to use a theorem that states that a structure in equilibrium has a minimum of internal potential (strain) energy. Inductance The storage of flow. Induction 1. The process by which flow produces potential without contact 2. Estimating the validity of a whole from observations of parts. Inertia The property of mass by which it retains its velocity when not acted on by a force. Ingeniarius The Latin root of the word engineer meaning someone who is ingenious in solving practical problems. Integrated circuit A complete electronic circuit manufactured as a single unit. Joule A unit of energy equal to the work done by 1 Newton moving 1 metre in the line of application of the force. It is 1 watt second. Material Having substance or matter that occupies space and has mass. Metal fatigue A deterioration of the mechanical properties of a material when subjected to a very large number of small repeated forces. Model A representation of an object. It may be physical, e.g. a three-dimensional (3D) representation of specific aspects of something such as an architectural model of a building. It may also be theoretical, e.g. a mathematical equation representing some physical process. It may be computational, e.g. computer representation of forces and displacements in a structure such as a finite element model or may be graphical such as a ‘walk through’ 3D pictorial representation of the interior of a building or an oil rig. Note that as models are representations of a reality, they are by definition incomplete and so describe the world only from a specific point of view. Newtonian physics Alternatively known as classical mechanics and based on the laws of motion formulated by Sir Isaac Newton (1642–1726). Newton One Newton is the force needed to accelerate one kilogram of mass at the rate of one metre per second squared in direction of that force. Object Anything we can think or talk about. It might be a material living thing such as an animal, plant or human being. It might be a material inanimate thing like a car or a bridge or it might be an immaterial and abstract idea like the number three or a belief in the afterlife or a fear of dogs.

216

Glossary

Object-oriented programming Programming a computer by creating ‘computer objects’ that interact with each other. These computer objects contain data as attributes and methods to process data. Phronesis Aristotle’s notion of practical wisdom and prudence. Potential Capable of becoming, voltage in electricity and velocity in mechanics. Power The capability of influencing, energy expended over time and measured in Watts. Practical wisdom A quality of discerning and judging—a way of looking at things with an ability to see the world in a coherent picture. It is the way a person constructs the world in which they operate, which in engineering is to do with having appropriate models to fit the situation. Progressive collapse A series of failures in a chain reaction such that the consequences are large compared to the initiating damage. Purpose An end, aim or goal or reason for why something is done or exists. Quantum computing Computers that use of quantum-mechanical phenomena to perform operations on data. They use quantum bits rather than binary digits. Radiation The process in which energy is emitted as particles or waves. Reading an object Observing, understanding and interpreting an object to give it meaning. Reductionism The philosophy that a complex system can be understood by splitting it down into its component parts and understanding them individually without reference to other parts or interactions between those parts. Relativity theory First developed by Albert Einstein (1879–1955) as a theory of physics based on two ideas. First that the laws of physics are the same for observers in uniform motion relative to each other and second that the velocity of light is the same regardless of that relative motion. Resilience The ability of a system to withstand or recover quickly from challenging conditions; to respond by detecting, preventing and, if necessary, handling disruptive challenges. It requires planning, learning, resources, watchfulness, coordination and cooperation. Resistance Opposition to flow through the dissipation of energy. Risk A combination of the chance of occurrence of an event in a context in the future. Router Software or hardware that transfers data between computers. SA Node The sinoatrial stimulus which depolarises heart muscle cells. Depolarisation is a change from the negative charge inside a cell (compared to the outside of the cell) to a positive one that allows electrical pulses to flow. Scalar A quantity that has magnitude only c.f. vector. Science That branch of knowledge that seeks to find truth. Semiconductor A material like silicon that conducts electricity but not as well as a good conductor like copper. Structure The difference between a random pile of component objects and a functioning object. For example, a building may have a structure of beams, columns, walls, etc.

Glossary

217

Sustainability Able to keep going over time. The ability to meet the needs of the time without compromising future needs. System An overused word with many meanings and uses. Used loosely it just refers to any group of connected objects. Systems thinking Systems thinking is a way of tackling problems with three important features—holons, connectivity and process. Technology There are four different uses of the word. To describe objects (such as a computer), as knowledge (we have the technology), as activity (crafting, inventing) and as an expression of human will (to make). Telegraph Long distance transmission of messages without the exchange of the physical object bearing the message. Transistor A semiconductor device that is a valve or an amplifier. Trebuchet A medieval war machine with a sling to hurl missiles. Truth The commonplace idea that there is a correspondence between a statement and the facts. But facts are true statements so there is a circular argument that can only be resolved by recognising that truth depends on context. Turbine A rotary engine where a flow of a fluid turns a shaft to drive a machine. Valve A switch or controller of flow. Vector A quantity that has both magnitude and direction c.f. scalar. VHF Very high frequency. Watts A unit of power equal to one joule per second. Work Exertion or effort and equal to one Newton being moved one metre or one Joule.

Index

A Academic versus vocational, 28, 29, 209 AC and DC, 111 Accidents-incubating, 45, 135, 187, 188, 203 Aerodynamics, 58, 83, 85, 86, 92 Aeroplanes -comet, 65 -solar impulse, 93, 94, 194 Aesthetics -in practice, 52 -relation with emotions, 7 Agricultural making, 156 Agriculture, 3, 111, 131, 153, 157, 160, 162, 197, 211 A learning odyssey, 32 All is not well on planet Earth, 9 Aristotle, 29, 34, 38, 182, 191 Architectural and structural form, 53 Architecture, 33, 46, 47, 51, 53, 57 Armstrong, William, 138, 139, 161, 207 Art -fine art, 3, 33, 47, 208 -functional art, 16, 24 -validation, 15 Artificial intelligence, 71, 102, 105, 106, 121, 123, 130, 195 Artificial Neural Networks (ANNs), 124 Art making, 15 As-if distinction, 19, 21, 24, 213

B Babbage, Charles, 105 Bardeen, John, 118 Bazalgette, Joseph, 3, 153, 168, 207

Ben’s story, 1 Bicycles -dandy horse, 74, 76 -penny-farthing, 69, 70, 74, 85 -safety, 69, 70, 79, 80, 85 Big data, 123, 124, 128–130, 178, 199 Big data becomes active, 123 Bioengineering, 6, 153, 198, 208 Blame, 3, 21, 22, 28, 43, 149, 208 Bramah, Joseph, 139, 166 Brattain, Walter, 118 Bridges -Menai, 21 -Morandi, Genoa, 4 -Pontcysyllte Aqueduct, 73 -Second Severn, 22 -Storebaelt, 59 -Tacoma Narrows, 59 -Tsing Ma, 59 -West Gate, 181, 182, 184–186, 188 Buildings, 3–5, 8, 15, 19, 31, 32, 38, 42–48, 51–53, 57–62, 64, 67, 101, 132, 133, 138, 146, 162, 164, 200, 204, 208 -Alfred P Murrah Federal, 60 -cathedrals, 5, 17, 31, 47, 182 -Grenfell Towers, 41–45, 52, 66, 204 -John Hancock, 62 -Ronan Point, 21, 41–44, 58, 60–62, 64, 66 -Sears (Willis) Tower, 59 -Shanghai World Financial Center, 62 -skyscrapers, 5, 50, 58, 60–62, 193 -WTC, New York, 4, 62, 64

© Springer Nature Switzerland AG 2020 D. Blockley, Creativity, Problem Solving, and Aesthetics in Engineering, https://doi.org/10.1007/978-3-030-38257-5

219

220 C Canals, 5, 16, 72, 73, 89, 96, 164 Cannon, 134, 136–138, 140, 145, 148 Capacitance, 164 Carnot, Sadi, 90 Cars, 1, 4, 6, 8, 13, 17, 20, 31, 33, 54, 58, 71, 74, 80, 85–88, 91–94, 99, 101, 115, 124, 127, 135, 140, 145, 148, 161, 199, 202, 203 Cathedrals, 5, 17, 31, 47, 182 Cave dwellers, 71 Cholera, 155, 163, 169, 170 Clerk Maxwell, James, 105, 112, 113 Climate change, 3, 9, 22, 27, 35, 93, 142, 149, 181, 184, 195, 209, 211 Collaboration, 31, 32, 35, 47, 53, 99, 142, 144, 156, 164, 178, 189–191, 201, 202, 206 Colt, Samuel, 138 Colt ‘peacemaker’ handgun, 138 Common enemy/common purpose, 7, 149, 183, 206 Complexity, 14, 15, 28, 30, 31, 37, 53, 64, 129, 178, 188, 208, 213 Complexity versus complicated, 31 Compressor, 98, 99, 100 Computers, 3–5, 8, 13, 16, 20, 22, 38, 56, 105–107, 117–125, 127–129, 136, 148, 164, 189, 190, 195–197, 200, 208, 215–217 Computing, 117 Concrete, 23, 42, 44–46, 50, 51, 53, 57, 58, 60–62, 67, 73, 146, 185, 190, 202 Context, 4, 7, 10, 19, 20, 21, 24, 31, 35, 37, 190, 192, 199, 201, 202, 203, 208 Cornish Man Engine, 182, 183, 210 Craft, 3, 10, 13–16, 19, 23, 24, 33, 34, 47, 135, 208, 213 Craft making, 14 Creativity, 5, 29, 33, 193, 199, 206 Creativity, problem solving and aesthetics -turning dreams into reality, 206 Crowden, James, 131, 149, 201 Cyber-attacks, 3, 135, 208 Cycling, 69, 70, 86

D Defence, 147 -policy, 138 Damping, 58 Decision making, 31, 35, 123, 148, 190, 201, 204, 205

Index Degrees of freedom, 56 Dependability, 19, 133, 148 Designing, 33, 52, 56, 90, 111, 133, 168, 182, 201 Disasters, 3, 21, 28, 30, 42, 43, 68, 135, 140, 149, 160, 181, 187 Disegno, 47 Dreaming, 8 Drinking water, 153, 162–164, 166, 197 Duty of care, 38, 52, 54, 127, 133, 134, 149, 192, 193, 208, 214 Dwelling, 10, 48, 67 Dynamo, 111, 113

E Early land transport, 71 Earthquakes, 30, 31, 54, 59, 60, 66, 67 Earth – the planet, 10, 58, 60, 144, 192, 209 Ecology, 9, 204 Economics, 4, 13, 38, 49, 138, 145, 147, 155, 160, 194, 202, 204, 205, 207, 209, 211, 212 Edison, Thomas, 111, 114, 207 Edith’s story, 69 Einthoven, Willem, 172–174 Electric motors, 74, 93 Electrocardiogram, 1, 172, 214 Electromagnetism, 102, 109 Electro-therapy, 171 Energy -conservation of, 88, 90 -electromagnetic, 105, 109 -flow, 18, 38, 70, 71, 100 -kinetic, 86, 95, 97, 100 -potential, 18, 55, 86, 88, 95, 96, 164 -strain, 55, 86, 88, 215 Engineering and technology, 23 Engineering as a general education, 22, 29, 33, 132, 154, 187, 201, 209 Engineering myths, 3, 28, 208 Engineering versus technology, 4, 23, 24, 129, 154, 198, 209, 210, 217 Engineers & engineering -as a people profession, 150 -definitions of, 5, 7 -fragmentation of, 51 -importance of, 7, 8 -in a bunker, 10, 207 -Italian artist, 33, 51 -making with science, 3, 19 -scope & responsibility, 8 -self-knowledge, 201, 212

Index -taken for granted, 53, 207 -types of, 3 Engineers at the heart of society, 201 Engines -electrical, 105 -heat, 86, 88, 90, 91, 94, 97, 112 -internal combustion, 33, 70, 74, 86, 91, 93, 100, 140, 161 -steam, 5, 9, 33, 67, 70, 86, 89–91, 100, 101, 159, 161 -turbine, 98, 208 Entropy, 90, 91 Equilibrium -dynamic, 87 -static, 55 -statically determinate, 55 -statically indeterminate, 55 Evolution, 130, 214 Extreme weather events, 9, 66, 149

F Failure, 3, 19, 21, 22, 30, 43, 60, 66, 156, 176, 181, 187–189, 197, 204, 208, 210, 216 Faraday, Michael, 3, 105, 109, 110, 112, 207 Fatigue -metal, 65, 68, 215 Feedback -positive, 58 Fessenden, Reginald, 115, 116 Fibres -glass, 120 -optical, 120 -plastic, 120 Field -gravitational, 112, 214 -magnetic, 109, 112, 129 -theory, 31, 112, 214 Fighting, 10, 129, 131 Fine art, 3, 33, 47, 208 Finite element method, 56, 214 Five principles, 28 Flourishing -human, 209 -personal, 182 Flourishing as achieving potential, 182 Fluids, 3, 49, 70, 86–88, 94–96, 101, 112, 161, 214, 217 -incompressible, 87, 95 -non-Newtonian, 87 Fluids and solids, 86 Flushing toilets, 153, 207

221 Flutter, 58, 59, 68, 214 Force -compression, 55 -of nature, 3 -shear, 55, 88 -tension, 50, 55 Form & function, 15, 17, 208 Foundations, 23, 24, 46, 48, 67, 73, 105, 190 Fragmentation, 51, 204 -of professions, 9, 51, 53, 181 Francis turbine, 88, 96, 97 Fuels -fossil, 89, 101, 195, 209 -renewable, 101 Fuel cells, 93, 194 Functional art, 16, 24 Functional art making, 16

G Generator, 18, 94, 95, 109, 111, 112, 164, 175, 176 Geometry, 19, 48, 87, 112, 192 Getting smaller, 129 Golden rule, 183, 210 Grand challenges, 3, 129, 160, 178, 181, 193, 194, 199, 202, 208, 212 Greatbtach, Wilson, 176 Grenfell Towers, 41–45, 52, 66, 204 Grounding Principles (PUPIL), 4, 28, 29, 207–209 Guiding mind, 45, 52 Gunpowder, 89, 134, 136–138, 148 Guns, 18, 134, 136, 138–140, 148

H Harbours -Mulberry, 146, 147 Hard versus soft systems, 190 Health -mental, 7, 196 Heart pacemaker, 1–3, 5, 6, 8, 153, 156, 171, 188, 197, 208 -evolution of, 171 Heat, 10, 70, 86, 88–91, 94, 97, 98, 100, 112, 129, 165, 215 Heat engines, 88 Holons, 189, 190, 198, 205, 217 Housing, 5, 43, 44, 57, 100 Human factors, 66

222 I Impedance, 215 Implantable Cardioverter Defibrillator (ICD), 2, 3 Incompleteness, 215 Incubating accidents model, 188 Independence, 23 Induction, 109, 215 Information -active, 122, 129 -passive, 122, 129 Ingenuity, 5, 22, 28, 29, 32, 33, 37, 44, 67, 74, 86, 88, 112, 135, 136, 149, 153, 157, 170, 171, 178, 182, 184, 189, 191, 193, 199, 200, 202, 205, 207–209 Instability, 59, 135, 147 Integrated circuit, 105, 116, 118, 129, 171, 215 Internal combustion engine, 33, 70, 74, 86, 91, 93, 100, 140 Internet, 8, 31, 34, 38, 119, 123, 128, 189, 208 Invention, 3, 13, 18, 24, 74, 105, 129, 165, 166, 177, 208

J Jacquard’s weaving loom, 107, 108 Jet engine, 31, 33, 98–101, 143 ‘Just war’, The, 134, 145, 149

K Kelpies, The, 15, 16 Kelvin, Lord – Thomson, William, 65, 112

L Land transport, 74 Langer, Robert, 153–155, 197 Law and economics as engineering, 209 Layers or levels of working, 210 Leadership, 22, 32, 128, 157, 205, 206, 210, 212 Learning, 1, 22, 28, 29, 31, 32, 35, 37, 43, 44, 53, 60, 66, 120, 124–127, 150, 182, 188, 193, 195, 198, 199, 201, 205, 208, 211, 216 Learning together requires leadership, 205 Life expectancy, 155 Literary engineering, 131

Index M Machines, 2, 6, 8, 29, 30, 33, 59, 69, 70, 72, 74, 76, 92, 95, 100, 105, 110–112, 117, 119, 120, 123, 124, 128, 136, 139, 140, 161, 177, 196, 197, 211, 217 Magnetism, 48, 102, 111, 112 Making, 3–10, 13–15, 21, 27, 31, 33–35, 42, 48, 49, 51, 69, 92, 95, 96, 101, 118, 123, 128, 136, 137, 148, 156, 160, 182, 190, 194, 196, 197, 199, 201, 204, 205, 207–209 Marconi, Guglielmo, 114, 115, 133 Medical engineering, 2, 9, 38, 56, 153, 156, 171, 178, 197, 198 Michelangelo, 33, 34, 47 Micro-electronics, 118 Military, 9, 22, 51, 90, 133, 134, 136, 138, 140, 145–148, 160, 184, 197, 213 Military targets, 145 Missiles, 3, 36, 132, 134, 136, 141, 142, 148, 208 Models, 16, 20, 21, 24, 66, 113, 124, 127, 128, 154, 176, 192, 199, 205, 206, 215, 216 Modulation -Amplitude (AM), 117, 213, 214 -Frequency (FM), 117, 213, 214 Moving, 10, 29, 32, 52, 65, 67, 70, 71, 78, 87–89, 92, 94, 98, 100, 118, 123, 129, 171, 172, 215 Moving on, 100 Mythos, 3, 20, 28, 30, 34, 39, 121, 163, 208 N National Grid, 111, 112, 164 Nature -laws of, 30, 31, 33 Needs and wants, 32 Negligence, 21, 38, 43, 188, 208 Neural nets, 124 Newton, Isaac, 19, 30, 48, 215 9/11, 4, 54, 61, 193, 203, 204 Nobel prize, 114, 118, 153, 154, 174 Nuclear weapons, 3, 9, 134–136, 141, 142, 148, 196, 208 O Objective, 15, 120, 121, 189, 190, 192, 208 Objects, 6, 7, 15–19, 31, 55, 70, 71, 119, 124, 129, 141, 164, 171, 200, 202, 214–217

Index Operating, 1, 111, 119, 144, 175 Opportunities, 27, 28, 32, 34, 48, 89, 128, 144, 189, 195, 199, 203, 209, 211 Otto cycle, 91 P Pacemaking, 171 Pantheon, Rome, 45, 46 Parsons, Charles, 97 Parts and wholes, 22, 28, 31, 71, 189, 209, 215 Part, Unintended, Preparedness, Ingenuity and Learning (PUPIL), 28, 29, 182, 208, 209 Passive and active information, 122 Pattern recognition, 105, 126 Patterns, 105 Patterns of electromagnetic energy, 105 Pelton wheel, 96 Perspectives, 122 Phronesis, 29, 191, 216 Plough -evolution of, 158 Pollution, 22, 35, 71, 89, 93, 101, 156, 165, 178, 194, 197 Polythene, 141 Popper, Karl, 35–37, 120, 121, 190 PQRST waves, 173, 174 Practical rigour, 192, 193 Practical wisdom, 28, 29, 33, 34, 67, 191, 193, 202, 209, 210, 216 Practical wisdom and rigour, 191 Prediction, 29, 31, 32, 36, 126, 203, 209 Preparedness, 28, 29, 32, 37, 43, 44, 52, 57, 64, 66, 149, 182, 189, 190, 193, 197, 205 Principles of practice, 27 Problem solving, 6, 14, 34, 36–38, 128, 132, 193, 206 Problem solving processes, 37 Problem solving with practical wisdom, 34 Professional silos, 208 Progressive collapse, 41–44, 60, 62, 66, 216 Progressive collapse and terrorism, 60 Properties -emergent, 43, 71, 188, 213–215 Public health, 3, 153, 155, 156, 166–168 Pugsley, Alfred, 42, 58, 59, 64–66 Punched cards, 117 PUPIL—see Grounding Principles, 4, 28, 29, 207–209 Purpose, 2, 4, 7, 15, 16, 19–21, 24, 35, 38, 56, 106, 125, 133, 146, 149, 159, 178,

223 183, 188–190, 195, 205, 206, 209, 214, 216

Q Quality, 2, 4, 7, 8, 14, 15, 17, 20, 24, 29, 34, 49, 57, 66, 117, 124, 155, 157, 164, 185, 187, 194, 199, 206, 211, 216 Quantum mechanics, Quantum computing, 30, 31, 39, 112, 129, 216 Queen Elizabeth Prize for Engineering, 154

R Radar, 44, 57, 117, 136, 140, 141, 147, 148 Radio, 18, 33, 105, 114, 115, 117, 140, 141, 207, 213, 214 ‘Reading’ a structure, 43, 53 Regulations -building, 41, 42, 44, 45, 67 Renaissance, 5, 14, 30, 32, 33, 46, 48, 62, 93, 207 Resilience, 29, 32, 193–195, 197, 198, 202– 205, 208, 210, 216 Resistance, 20, 51, 54, 78, 84, 101, 103, 115, 150, 164, 215, 216 Revisiting patterns, 188 Rigour -mathematical, 192 -practical, 192, 193 Risk, 2, 9, 10, 22, 34, 42, 45, 57, 67, 69, 88, 120, 123, 129, 130, 135, 136, 162, 165, 188, 190, 194, 198, 200, 202, 203, 208, 216 Robots, 9, 53, 102, 127, 128, 133, 148, 199 Robustness, 29, 204 Rockets, 101, 134, 143, 144, 148, 149, 207 Rockets and missiles, 143 Ronan point, 21, 41–44, 58, 60–62, 64, 66 Ronan point and Grenfell towers, 41 Royal Academy of Engineers, 5, 34, 154

S Safety, 9, 21, 44, 45, 52, 54, 62–70, 79, 80, 85, 97, 109, 127, 134, 144, 159, 161, 187, 196, 203, 209 Salginatobel bridge, 17 Scanners, 8, 31, 106, 156, 207 Science, 3, 5–7, 9, 10, 14, 19–21, 23, 24, 27– 31, 33–35, 48, 58, 67, 68, 90, 112, 113, 121, 128, 130, 133, 134, 136, 141, 143, 144, 192, 198, 201, 202, 207, 208, 210, 212, 214, 216

224 -incompleteness, 28 Semantics and pragmatics, 106, 120 Sewers, 165, 166, 168–171 Shelter, 45 Shockley, William, 118 Signalling, 109, 123 Simpson, Christopher, 181, 186 Social media, 3, 9, 22, 31, 122, 123, 129, 135, 193, 195 Solar cells, 31, 93, 194, 199 Solar Impulse aeroplane, 93, 94, 194 Solids & fluids, 3, 70, 86–88, 101 Space, 21, 22, 24, 53, 57, 70, 87, 101, 106, 107, 112, 126, 134, 143, 144, 149, 174, 182, 196, 200, 207, 213–215 Spark gap transmitter, 115, 140 Steam engine, 5, 9, 33, 67, 70, 86, 89–91, 100, 101, 161 STEMM, 24, 25 Structural safety, 62, 64–66, 68 Subjective, 120, 121, 189 Sustainability, 29, 57, 149, 160, 193, 194, 197, 198, 204, 211, 217 Sustainability versus war, 149 Sustainable development goals, 194, 211 Syntax, 105–107, 115, 120, 121, 195, 199 Systems -biology, 31, 39 -engineering, 31, 140, 148, 178, 188 -thinking, 189, 190, 197, 198, 206, 208, 209, 217

T Targets, 145 Technology triumphalism, 34, 136 Telecommunications, 8, 105, 120, 128, 145 Telegraph, 33, 109, 114, 115, 217 Terrorists, 3, 4, 54, 60, 62, 134, 148, 196, 208 Thermodynamics, 33, 67, 90, 91 -first law of, 90, 91 -second law of, 91 Things, 2–9, 13–15, 21, 24, 28, 31, 34, 35, 58, 64, 66, 67, 70, 71, 89, 113, 120, 121, 123, 133, 170, 188–190, 193, 198, 199, 203, 206–210, 215, 216 Tractors, 161 Transistor, 105, 118, 129, 171, 176, 199, 217 Transmitter -radio, 105, 115, 141 -spark, 115, 140 -telegraph, 115, 217

Index Transmitting the human voice, 115 Transport -air, 71 -land, 71, 74 -water, 72 Trebuchet, 134, 136, 137, 148, 217 -early ingenuity, 136 Trust, 127, 149, 165, 193, 202, 204, 206, 210 Turbine, 67, 86, 94–100, 112, 164, 208, 217 -Francis, 88, 96, 97 -gas, 95, 97–100, 208 -impulse, 95, 96 -jet, 86 -Kaplan, 97 -Parsons, 97 -Pelton wheel, 96 -reaction, 95, 97 -waterwheel, 94, 95 -windmill, 94, 95, 97 Turner, Barry, 43, 45, 135, 187, 188, 203, 209

U Uncertainty -fuzzy, 32 -incompleteness, 24, 198 -known unknowns, 135 -randomness, 30 -unknown unknowns, 148, 182, 198 Unintended consequences, 22, 28, 29, 31, 32, 35, 37, 42–44, 71, 102, 123, 129, 136, 148, 149, 178, 182, 190, 193, 195, 196, 206, 208, 209 United Nations (UN), 27, 38, 150, 194, 209 Universe -clockwork, 29, 30, 34 Unmanned Aerial Vehicles (UAV), 147, 148

V Vacuum tube, 115, 116, 118 Valves, 89, 118, 129, 139, 164, 175 Vehicles -electric, 70, 93, 101 -hybrid, 93 Venturi effect, 87, 88 Vibration, 43, 59, 115 Virtual reality, 196 Vitruvius, 15, 47, 208 Volts, 70, 106, 118 Vulnerability, 45, 59, 60, 134, 147, 188, 204

Index W Wagons, 5, 10, 71, 74 War, 132, 134 -Crimean, 109 -just–bellum justum, 134, 149 Warfare, 3, 10, 32, 36, 38, 72, 133, 134, 135, 136, 140, 141, 148, 196 Waste, 22, 73, 89, 141, 153, 155, 156, 163, 165, 168, 177, 197, 200, 207 Waste disposal, 165, 168 Water transport, 72 Waterworks, 163, 164 Waves -gamma, 114 -light, 112, 114, 115, 129, 216 -radio, 114, 141 -ultraviolet, 114 -X ray, 114 Weapons, 135

225 -drones, 3, 134–136, 147, 148, 196, 197, 208 -nuclear, 3, 9, 134–136, 141, 142, 148, 196, 208 -missiles, 3, 134, 136, 141, 142, 148, 208 Weaving, 105, 107, 108 Well-being, 155 West Gate Bridge, 181, 184, 186, 210 Wheel, 15, 69, 71, 74, 85, 86, 91, 94–96, 138, 139, 157, 161, 203 Whittle, Frank, 3, 98, 207 Why, how, who, what, where, when, 190 Why war, 134 Will this happen again, 187, 188, 208 Wind, 16, 18, 36, 42–44, 54, 58, 59, 66, 68, 72, 88, 94, 95, 97, 101, 157 Wind and vibrations, 58 Women in AI, 128