The Canadian Light Source: A Story of Scientific Collaboration 9781487537470

This book details the people and politics involved in the development of the Canadian Light Source, the benefits to be g

170 100 5MB

English Pages 184 Year 2020

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

The Canadian Light Source: A Story of Scientific Collaboration
 9781487537470

Citation preview

THE CANADIAN LIGHT SOURCE

The Canadian Light Source A Story of Scientific Collaboration

G.M. BANCROFT WITH D.D. JOHNSON

UNIVERSITY OF TORONTO PRESS Toronto Buffalo London

©  University of Toronto Press 2020 Toronto Buffalo London utorontopress.com Printed in Canada ISBN 978-1-4875-0806-7 (cloth) ISBN 978-1-4875-3748-7 (EPUB) ISBN 978-1-4875-3747-0 (PDF)

Library and Archives Canada Cataloguing in Publication Title: The Canadian Light Source : a story of scientific collaboration / G.M. Bancroft   with D.D. Johnson. Names: Bancroft, G. M. (George Michael), 1942– author. | Johnson,   D. D. (Dennis Duane), 1938– author. Description: Includes bibliographical references and index. Identifiers: Canadiana (print) 2020020758X | Canadiana (ebook) 20200207644 | ISBN   9781487508067 (hardcover) | ISBN 9781487537487 (EPUB) | ISBN 9781487537470  (PDF) Subjects: LCSH: Canadian Light Source – History. | LCSH: Synchrotrons – Research –   Saskatchewan – Saskatoon – History. Classification: LCC QC787.S9 B36 2020 | DDC 539.7/35 – dc23

This book has been published with the generous assistance of the Canadian Light Source Inc. and the University of Saskatchewan. University of Toronto Press acknowledges the financial assistance to its publishing program of the Canada Council for the Arts and the Ontario Arts Council, an agency of the Government of Ontario.

Contents

List of Figures  vi Acknowledgments  vii List of Acronyms and Abbreviations  ix 1 Introduction  1 2 The University of Saskatchewan: The Electron Accelerator, Technical and Engineering Expertise, 1930s–1990  8 3 The University of Western Ontario: The Beamline and Experimental Expertise, 1970s–1990   21 4 Formation of the Canadian Institute for Synchrotron Radiation and Competition between Western and USask, 1989–97  42 5  The Creation of the Canada Foundation for Innovation  73 6  My Role as Interim Director, 1999–2001  87 7  The CFI: Goals, Impact, and Paul Martin  105 8  The Positive Impact on USask and Canadian Science  111 Appendix 1 Synchrotron Facilities and Synchrotron Science: A Brief Overview  143 Appendix 2 Canadian Institute for Synchrotron Radiation: Announcement of CFI Funding, 1999  149 References  153 Index  159 About the Authors  169

Figures

  1   2   3   4   5   6   7   8   9 10 11 A1 A2

Sod-turning for SAL, May 10, 1962  13 Leon Katz at the new LINAC at SAL, 1964  14 Les Dallin and Mark Silzer, ca. 1990  17 Dennis Skopik outside the front entrance to SAL, ca. 1990  18 Doug Richardson and Dennis Johnson at SAL, 1999  20 The official opening ceremony of the first CSRF beamline (the Grazing incidence “Grasshopper” beamline) at the Tantalus synchrotron in early 1983  33 The official opening of the third CSRF beamline (the SGM beamline) in 1999 37 Some of the important Canadian synchrotron scientists in the quest for CLS 39 CLS employees trained at the Aladdin synchrotron standing with the CSRF Grasshopper monochromator at CLS  40 London Free Press cartoons, May 23 and June 5, 1996  62 The official opening ceremony of CLS in September 2001  103 Schematic of the Aladdin synchrotron at the University of Wisconsin 144 Schematic of a hard x-ray beamline  145

Acknowledgments

I would first like to thank the Canadian Light Source for inviting me to give a talk on the history of CLS – entitled “Towards the CLS: Teamwork and P ­ olitics” – to the 18th Annual CLS User’s meeting in May 2015. This talk greatly encouraged me to address some of the older history of CLS, and Sandra Ribeiro and Les Dallin kindly sent me five photos that I incorporated into my presentation (and which comprise the first five figures in this book). That experience inspired me to write the larger history of synchrotron radiation in Canada. Dennis Johnson had a similar idea, and we decided to combine our knowledge into this book. What follows is meant to thank the scores of dedicated people, both named and unnamed, that contributed to CLS mainly up to 2001. This book could not have been written and completed without a lot of help from many people. Specifically, talks with Les Dallin, the late Jack Bergstrom, and Johannes Vogt were very helpful for the early sections on the Saskatchewan Accelerator Laboratory. Mark Boland and T.K. Sham set me straight on fourth-generation sources. Many others have contributed data and comments and suggestions for new sections. Brian Yates and Yongfeng Hu suggested that I write about the outstanding research at CLS (covered in chapter 8) and made many other useful comments. Sandra Rebeiro, Jeff Warner, Jeff Cutler, Beryl LePage, Lavina Carter, and Ning Chen from CLS provided useful data and comments. Greg Fowler, John Tse, and Tim Kelly from USask, Wayne Nesbitt, Phil Dean, and T.K. Sham from Western, and Adam Hitchcock from McMaster all provided useful comments and suggestions. My son David Bancroft gave useful comments on the whole manuscript. Nils Petersen and my editor at the University of Toronto Press, Stephen Jones, have provided many very useful suggestions as well, which hopefully have led to a more balanced, focused manuscript. Daryl Crozier, Jeff Cutler, Betty Harper, Adam Hitchcock, Yongfeng Hu, Masoud Kasrai, Sandra Ribeiro, T.K. Sham, and Brian Yates were very helpful in retrieving high-quality photos/diagrams for the figures. We are grateful to the above people, and of course to everyone discussed in the book (and the whole CLS staff who have not been named) who contributed to this fascinating story.

Acronyms and Abbreviations

All the acronyms and abbreviations have been spelled out when they are first introduced. However, it is useful to have a list to which readers can easily refer as they make their way through the text. Note that it is normal to drop the definite article in front of many of these acronyms. For example, “the National Research Council” becomes simply “NRC.” AECL ALS APS BDI BPAC CAP CCP CERC CFI CIHR CII CISR CLS CLS Inc. CMRF CNSC CRC CSC CSRF DCM EROS ESCA

Atomic Energy of Canada Ltd. Advanced Light Source, Berkeley Advanced Photon Source, Chicago Beamline Development Inc. Beamline Planning and Access Committee Canadian Association of Physicists Centre for Chemical Physics, Western Canada Excellence Research Chair Canada Foundation for Innovation Canadian Institutes of Health Research Canadian Isotope Innovations Inc. Canadian Institute for Synchrotron Radiation Canadian Light Source Canadian Light Source Inc. Committee on Materials Research Facilities Canadian Nuclear Safety Commission Canada Research Chair Canadian Society for Chemistry Canadian Synchrotron Radiation Facility, Madison, Wisc. Double Crystal Monochromator Electron Ring of Saskatchewan Electron Spectroscopy for Chemical Analysis

x  Acronyms and Abbreviations

ESRF European Synchrotron Radiation Facility, Grenoble EXAFS Extended X-ray Absorption Fine Structure FAC Facility Advisory Committee ISW Interface Science Western LINAC Linear Accelerator MAC (1) Machine Advisory Committee, CLS MAC (2) Multi-disciplinary Assessment Committee, CFI MBA Multi-Bend Achromat MRC Medical Research Council NRC National Research Council NRCan Natural Resources Canada NRU National Research Universal Reactor, Chalk River NSERC Natural Sciences and Engineering Research Council NSLS National Synchrotron Light Source, Brookhaven OC Organic Carbon OIT Ontario Innovation Trust OSC Ontario Synchrotron Consortium PMAC Pharmaceutical Manufacturing Assoc. of Canada PSL Physical Sciences Laboratory, Univ. of Wisconsin PSR Pulse Stretcher Ring RF Radiofrequency ROC Review Oversight Committee SAC Science Advisory Committee SAL Saskatchewan Accelerator Laboratory SAXS Small Angle X-ray Scattering SECD Sask. Economic and Cooperative Development SED Space Engineering Division, USask Res. Park SGM Spherical Grating Monochromator SMEs Small and Medium Sized Enterprises SR Synchrotron Radiation SRC Synchrotron Radiation Centre, Madison SSHRC Social Sciences and Humanities Research Council SSI Saskatchewan Synchrotron Institute SSW Surface Science Western STXM Scanning Transmission X-ray Microscopy TRIUMF Tri-University Meson Facility UAC Users Advisory Committee UHV Ultra-High Vacuum UMA Underwood, McLellan and Associates UPS Ultraviolet Photoelectron Spectroscopy USask University of Saskatchewan VIDO Vaccine and Infectious Disease Organization

Acronyms and Abbreviations  xi

WED Western XANES XAS XEOL XFEL XPS

Western Economic Diversification University of Western Ontario X-ray Absorption Near Edge Structure X-ray Absorption Spectroscopy X-ray Excited Optical Luminescence X-ray Free Electron Laser X-ray Photoelectron Spectroscopy

1 Introduction

The Canadian Light Source (CLS) in Saskatoon, the largest Canadian s­ cience project of the last fifty years, resulted from the concerted efforts of dozens of scientists, university administrators, post-docs, and graduate students from across Canada, as well as Canadian government (federal, provincial, and municipal) and private industry employees. In addition, scientists and o ­ rganizations from many countries around the world provided essential expertise and ­co-operation. Indeed, this story is a wonderful example of scientific co-operation and international “free trade,” which happens so often in science: it was a “win-win” for a huge number of scientists, engineers, and medical researchers at the University of Saskatchewan (USask), the city of Saskatoon, the province of Saskatchewan, and indeed the whole of Canada. Such collaboration and co-operation, essential in nearly all areas of human endeavour for the greatest overall benefit, is the major theme of this book. The account given here can be broken into two periods: (1) the 1930s to 1972, and (2) 1972 to 2001, with a later section on the subsequent success of CLS. Neither myself nor Dennis Johnson was involved in the first period, so we have relied on works of local history to describe some of the critical events in chapter 2. For the second period, we have deliberately given a very personal ­account of some of the events in which we were intensely involved. This personal ­approach is not used in the vast majority of scientific articles and ­scientific-historical writings. We think that it is important to demonstrate that scientists and ­administrators have all of the personal characteristics – both good and bad – of human beings in any venture. Fortunately, in this case the positive human characteristics dominated: for example, dedication, co-operation, altruism and generosity, excitement, joy, and humour. But ­inevitably in this long saga, negative human characteristics appear a number of times – for example, apathy, ignorance, frustration, and greed. Thus, much of our account of the 1972–2001 period focuses not on the science, but rather on the human interactions that led to CLS – what we might term the

2  The Canadian Light Source

s­ociology of CLS’s development. We hope that this personal approach will generate more interest from administrators and academics in all disciplines, as well as the scientific layman. A somewhat amusing story from my twenty-five-year odyssey to fund a ­Canadian synchrotron at the University of Western Ontario (henceforth ­Western) is probably a good metaphor for this book. In early January 1997, my wife Joan drove me to the London, Ontario, airport to catch a flight to Toronto, where I was to meet my USask colleagues Dennis Skopik and Dennis Johnson. We had planned to spend three days visiting universities and federal politicians, such as Dennis Mills in Toronto, to gain support (both scientific and financial) for the CLS project, which had been awarded to USask but was not yet funded. When I got to the airport in London, I noticed that I had two different types of shoes on – one brown and one black. This was rather ­embarrassing (especially since this was not the first time!). But my wife told me, “just keep moving.” On that trip I had no shortage of trouble giving talks on one leg, trying to cover one shoe with the other. We had a lot of laughs over those shoes! That “just keep moving” motto symbolizes many of the fundamental a­ spects of my (and several other academics and administrators) t­wenty-five-year quest for a Canadian synchrotron. Most importantly, I wish to stress here the co-operation and encouragement from not only my wife but at least a ­hundred students, post-docs, faculty, staff, administrators, and lawyers from Canada and many other countries around the world. Also, it is important to mention the mistakes and many rejections; the dogged ­determination and perseverance necessary to keep a high-risk goal in sight for years and to “keep moving” it forward; the huge effort from many at USask; and the constant challenge, ­enjoyment, and joy of scientific research often mixed with a good deal of ­humour. This account of the building of the synchrotron also includes several other themes: the unselfishness of many university and NRC, NSERC, and CFI council ­academics and administrators from Canada and abroad who go out of their way to help colleagues with absolutely no extra remuneration; the intense competition for funding within universities (e.g., humanities versus sciences) in ­Canada and abroad; the critical importance of taking big risks for big ­rewards; the conflicts ­between individuals for funding; the conflicts ­between groups (e.g., university versus provincial authorities) who sometimes want to control the project, thereby benefiting financially and/or academically; the dramatic differences in p ­ roductivity and effectiveness (sometimes by an ­order of magnitude or more!) of individual scientists and administrators who are f­ ormally equally qualified; the fortuitous convergence of outstandingly p ­ roductive ­individuals; the t­ iming of funding; and, finally, the importance of academic research for economic development.

Introduction 3

a)  The Synchrotron: Intense Beams of Light A modern synchrotron facility such as CLS produces twenty to thirty tiny, very intense beams of light from the far infrared to the hard x-ray region (a large part of the electromagnetic spectrum). These beams of light have become ­essential worldwide for detailed research and analysis of matter in most scientific ­areas (e.g., biology and medicine; advanced materials in chemistry and physics; e­ nvironment and mining; and agriculture). The beams are produced by a large circular electron accelerator called a synchrotron or storage ring (see appendix 1, figure A1). The electrons in this “ring” are accelerated to close to the speed of light. When these high-energy electrons are bent around the ring by large magnets, they produce from twenty to thirty light beams, tangential to the ring. The intensities (or brightness) of these beams can be up to a billion times the intensity of light from the sun. Beamlines are constructed to utilize these beams in the infrared, ultraviolet, and x-ray regions. At the end of each beamline, experimental chambers are constructed with various detectors for specific types of analysis (see ­appendix 1, figure A2). The entire CLS facility of accelerators, beamlines, and experimental chambers is larger than a football field. All the twenty to thirty beamlines can be utilized at the same time, and the facility operates seven days a week, ­twenty-four hours a day for most of every year (a total of over five thousand hours). Over a thousand researchers a year from most parts of Canada and many other countries use CLS for important research in most scientific areas. Such a facility can have a very large scientific impact. Most of the experiments at CLS utilize both existing and novel spectroscopies in the infrared and x-ray regions. Also, diffraction, scattering, and imaging are prominent. These techniques examine the interaction of light with matter, usually with a variable energy monochromatic source impinging on a sample. The amount of light absorbed at each energy is recorded, and a spectrum or image is produced that can be used to characterize that sample. A synchrotron is an ideal source of light for spectroscopy, diffraction, and imaging, mainly because the light beams are very small, very intense, and monochromatic. The beams are particularly useful for unique chemical micro-imaging of most materials. b)  The Scope of the Book Gerry Kline1 published an informative article on the history of CLS in the ­October 20, 2004 issue of the Saskatoon Star Phoenix. I have also previously documented some details of the Western synchrotron effort.2–4 This book, however, has a much broader scope, one that emphasizes the seventy-year ­commitment at USask5–17 and the human element, with the aim of making it more interesting to a larger audience, including non-scientists. Indeed, the

4  The Canadian Light Source

many positive and negative human resonances were absolutely critical to the successful creation and operation of the largest Canadian scientific facility in the last fifty years. i)  The Early History at USask and Western until 1989 The history of CLS begins with the establishment at USask of three fundamental research programs over seventy years ago: the spectroscopic research of ­Gerhard Herzberg in the 1930s; the cancer/therapy program of Ertle H ­ arrington in the 1930s; and the purchase of the first electron accelerator (the betatron – the first electron accelerator in Canada) by Ertle Harrington, Harold Johns, and Leon Katz of USask in 1946. This acclerator was used for cancer therapy (headed by Harold Johns) and for the nuclear physics program (headed by Leon Katz). In 1962, Katz and his colleagues then developed a very strong ­nuclear physics program with the aid of new linear and “circular” electron accelerators, which lasted into the 1990s. This accelerator expertise was essential for the construction of the CLS synchrotrons. These developments are described in chapter 2. Equally important contributions to CLS’s development (described in ­chapter 3) came later in the 1970s from Western. Bill McGowan (chair of the physics department and director of the Centre for Chemical Physics [CCP] at Western) was the first Canadian to recognize the broad scientific potential of synchrotron radiation (SR). He proposed the first Canadian synchrotron at Western in 1973 after the first SR experiments worldwide in the late 1960s. ­McGowan followed up with a 1976 application to the National Research C ­ ouncil (NRC) for a $10 million synchrotron at Western. With no Canadian users at the time, this application was rejected. I then followed up with a successful beamline application to NRC (called the Canadian Synchrotron ­Radiation Facility) to be constructed at the synchrotron outside Madison, Wisconsin in 1979. Two more beamlines at the facility followed in the next twenty years (along with Canadian participation at other US synchrotrons), enabling excellent research by many Canadian users. This facility also developed the essential beamline technical and scientific expertise for a future Canadian synchrotron. ii)  The Push for CLS, 1989–2001 Chapters 4 and 5 describe the concentrated efforts of many Canadian scientists, university administration, lawyers, and government officials to obtain scientific approval and funding for CLS. Some of these elements will be common to any large project, although the funding and political complexities in Canada appear to be much greater than in most countries. This section begins in chapter 4 with the creation of the Canadian Institute for Synchrotron Radiation (CISR) in 1989, followed by the development of the large, multidisciplinary scientific

Introduction 5

community required for such an expensive project, and the struggle for funding and control. The positive role played by the Natural Sciences and Engineering Research Council (NSERC) in the project’s development is emphasized. This is followed by the intense competition (conducted by NSERC) between USask (directed by Dennis Skopik) and Western (directed by myself) to land the facility, and the creation of a new scientific organization by the Liberal government of Jean Chrétien, the Canada Foundation for Innovation, that funded 40 per cent of the initial project at USask. A myriad of technical, scientific, and complex political hurdles had to be overcome, and although most of the 60 per cent matching funds had been raised from nine different organizations, the project was still short $32 million. Also, the management structure was still not in place. Even after the approval and partial funding of the endeavour in 1996–7, several years of discussion and controversy followed before the funding ­ and management structure were established at USask in 2001. In chapter 6, I ­describe my role as interim director from 1999 to 2001, again emphasizing the multiple challenges inherent in raising $19 million, mainly from Ontario and Alberta, for a project in Saskatchewan, along with establishing the management structure and the guidelines for funders and users. By the time I left Saskatoon in October 2001, the budgeted $140.9 million had been raised, and the management structure was in place. The important role played by the Canada Foundation for Innovation (CFI) in CLS and Canadian science more broadly is described in chapter 7. The lobbying by Canadian scientists, administrators, and lawyers to fund CLS helped Paul Martin and the Liberal government create the CFI; and CLS was, in turn, a strong influence on the CFI’s national focus. iii)  The Impact of CLS since 2005 In chapter 8, the very positive impact of CLS on USask and Canadian science is described. CLS, which began operating in Saskatoon at the end of 2004 and started accepting users in 2005, is now a thriving operation with over 250 ­employees from 20 countries, an average of over 280 people in the building on any given day, operating grants of over $30 million per year, equipment grants of over $300 million, over 1,000 users per year, major industrial interactions, and the formation of a spin-off company, Canadian Isotope Innovations (CII), which produces medical isotopes. Even before it began operating, Peter MacKinnon, the president of ­USask, declared in the university’s 2000 annual report that “CLS symbolizes the ­future of this university.” In the last section of this book, we show that CLS has ­indeed made a dramatic positive impact on USask. For example, from 1999 to 2016, USask had the largest per cent increase in research funding of any major ­Canadian university, largely due to CLS funding. CLS, along with other major

6  The Canadian Light Source

projects such as the Vaccine and Infectious Disease Organization (VIDO), has enabled USask to take its place as a major Canadian research university. Research at CLS has been focused on four themes: advanced materials, health, environment, and agriculture. Some of the important, much-cited research from these areas is described briefly for the layman. Interdisciplinary and international co-operation is shown to be critical for much of this research. Given the difficult politics of funding at both USask and Western (indeed, at all universities!), it is probably amazing that this facility was actually built at any university in Canada. Moreover, it was built on time and on budget for $140 million – an outcome that was the product of a large number of talented, dedicated, and unselfish scientists, engineers, and administrators. c)  The Goals and Objectives of This Book The major goal of this book is to provide the definitive history of CLS’s development from the 1930s to 2001.* The financial, research, and educational success of CLS after 2001 and up to 2018 is also described in less detail. Instead, we focus on the preceding decades of work by dozens of dedicated Canadians to develop the technical and scientific expertise, mainly at USask and Western; the intense lobbying of administrators and politicians with many potential funders over many years; and the difficult political challenges at USask and Western, as well as with the province of Saskatchewan. Another important goal of this book is to show that the Canadian method of funding the capital and operating costs for such a large facility is extremely complex. For example, it does not seem reasonable that thirteen different ­organizations were required to provide the $140.9 million for CLS initially with at least eighteen separate contracts; and it is certainly not reasonable that the operating funding has come from six or more different organizations, all with different regulations and requirements. Canada must develop better mechanisms for funding large and important scientific facilities in the near future. We have several other goals that are not normally addressed in scientific articles or books. First, it is important to acknowledge and celebrate the contributions of dozens of people in this saga, as well as the contributions of many different universities and government organizations such as NSERC and the CFI. Second, it is critical to emphasize and celebrate the international * This account does not include the other Canadian synchrotron beamline initiatives at US facilities from the mid- to late 1990s by other Canadian researchers – namely Daryl Crozier (Simon Fraser University), Mark Sutton (McGill University), and Adam Hitchcock (­McMaster University) at the Advanced Photon Source (APS) in Chicago (Crozier, Sutton) and the ­Advanced Light Source (ALS) at Berkeley California (Hitchcock). Crozier’s beamline became part of CLS in 2015.

Introduction 7

collaboration and co-operation engendered in this project and most academic research today – such collaborations are becoming an essential part of solving ever larger global problems such as the migrant crisis and climate change. Third, it seems important to give glimpses of humanity throughout this book, including instances of humour and altruism, along with a few comments on personal interests such as sports, music groups, nature, and religion. Each of these plays a role in maintaining a healthy and positive attitude, especially in stressful times. Team sports (such as rowing and curling) are a wonderful way of developing teamwork, interpersonal relationships, and “fair” competition; and music groups (such as choirs and bands) are also a great way of developing teamwork and compassion toward others. Certainly, some of these activities and interests demonstrate that losing is common; the main thing is that everyone puts in their best effort, win or lose. This book shows that dozens of people put in their best effort! Although a good part of this story chronicles my personal odyssey with synchrotron radiation, Dennis Johnson has provided a good deal of the material for chapters 4 and 5, as well as some of the early history in chapter 2. He has also carefully critiqued the whole manuscript several times.

2 The University of Saskatchewan: The Electron Accelerator, Technical and Engineering Expertise, 1930s–1990

a)  The Pre-Accelerator Days, early 1930s–1946 It seems appropriate to begin this story at the University of Saskatchewan (USask) in 1933.* A young academic in the chemistry department, John Spinks,† spent a sabbatical working with a young spectroscopist, Gerhard Herzberg, in Germany during the 1933–4 academic year. Spinks was asked to take the leave from USask with little pay because of the dire financial situation at the university, which was then in the grips of the Great Depression. Spinks and Herzberg roomed in the same house, and they worked very hard to get their results published in eight papers. After this year in Germany, Spinks, knowing that Herzberg could not stay in Germany because Hitler had passed the Law for the Restoration of the Career Civil Service – which required all Jews in government service to be fired – convinced USask to offer Herzberg a guest professorship in the physics department. Even though Herzberg was not a Jew, he was denied the right to teach in Germany because his wife, Luise, was a Jew. Herzberg searched all over the Western world to find an academic position, but only USask came up with an offer. Spinks made it quite clear in a letter to ­Herzberg “that Saskatoon was a long way from anywhere, and that visits to other research centers in Canada and the U.S. were likely to be few and far between.”5 As an aside, “the Jews made up less than 1% of the German population of 65 million, most had lived in Germany for several generations, they had fought in wars, and contributed to its renowned culture.”5 This terrible abuse of human rights by perhaps the world’s most educated and cultured population, with

* Much of this material is detailed in the Herzberg biography by Boris Stoicheff,5 who also ­appears later in this book. † Much later, in 1959, Spinks became the president of USask, a position he held for fifteen years.

The University of Saskatchewan  9

the much more incredible evil that followed, was to play an important role in the future development of USask: it pushed the school to hire one of the most brilliant and productive scientists in the world.* As we will see in the ensuing pages, it could be argued that the meeting of Spinks and Herzberg, along with Herzberg’s hiring, were the most important events in eventually landing the Canadian Light Source (CLS) over sixty years later. Herzberg’s position at USask was funded for two years with a $4,500 grant from the Carnegie Foundation, and because of his outstanding work he was soon made research professor at USask while building a spectrograph for $1,500, which was paid for by the American Philosophical Society. He did outstanding work at USask, and really put science on the map there until 1945. He was recruited by the University of Chicago, where he did a brief stint before coming back to Canada to work for the National Research Council (NRC) in Ottawa as principal research scientist and director of the Physics Division in 1948. He pioneered infrared spectroscopy at USask, and indeed in Canada and worldwide. His first book, Atomic Spectra and Atomic Structure, published by the University of Saskatchewan in 1944,6 was a classic for generations of chemistry and physics students. This book was translated from German by none other than John Spinks. Herzberg’s later, very large books have been tremendously important to a large fraction of chemistry and physics students – indeed, they are legendary in the scientific community, and led to his being awarded the ­Nobel Prize in Chemistry in 1971. The first electron accelerator expert, Leon Katz, was hired by the USask physics department in 1946 as a formal replacement for Herzberg, and Katz was heavily involved in the first accelerator purchase by Harold Johns and colleagues. Of equal importance to the development of the accelerator and nuclear physics program at USask was the career of Ertle Harrington, a USask physics professor. He built and operated a radioactive radon (Rn) plant from 1931 to 1962, extracting the radioactive gas emitted by Rn in solution.7 Dr. Harrington distributed Rn to both of the Saskatchewan cancer clinics in small glass tubes, or “seeds,” for radiotherapy using the high-energy gamma (denoted γ) rays from radioactive Rn with a 3.5 day half-life. In 1945, Dr. Allan Blair, a radiotherapist and the director of the Regina Cancer Clinic, wrote Dr. Harrington suggesting that “a full time physicist be hired jointly by the Saskatchewan Cancer Commission and the University.”7 “Six days later Harrington replied offering full co-operation.” In March 1946, Harrington * Like most civilized human beings, I have struggled mightily with the Nazi atrocities nearly all my life, especially after my father (who grew up in Regina and lived in Saskatoon from 1935 to 1942, where I was born in 1942) emphasized the plight of the Jewish people in the 1930s and 1940s to me when I was twelve or thirteen.

10  The Canadian Light Source

hired Harold Johns, who became famous for leading the development of radiation sources for cancer therapy. b) The First Electron Accelerators, 1946–65, and the Early Conflict between the Humanities/Social Sciences and the Basic Sciences In his early travels in North America to study “modern” concepts of cancer therapy, Johns heard a visitor from England suggest that both gamma rays from radioactive cobalt (60Co) and gamma rays from electron bombardment of a heavy metal such as tungsten (W) might provide suitable radiation sources for radiation therapy. Upon returning to USask in 1946, Johns, with Blair’s support, initially asked for a betatron electron accelerator to yield 25 MeV (MeV = million electron volts) electrons – the highest energy electrons available at that time. The president of USask, James S. Thompson, in turn talked to Dr. C.J. MacKenzie, head of NRC, about funding. The cost of a betatron was projected to be $80,000, a large sum in 1946. The Atomic Energy Control Board of Canada provided $30,000 and the Saskatchewan government funded the heavily shielded concrete enclosure (seven metres thick in some places) in the physics building to house the betatron. The magnitude of such a project – big engineering, new science, high vacuum, high cost, etc. – was already evident. This was pioneering in Canadian science, and in a relatively poor province at that. In May 1948, Harold Johns, R.N. Haslam, and Leon Katz travelled to ­Milwaukee, Wisconsin to see their new betatron at Allis Chalmers, the agricultural equipment manufacturer. The 25 MeV betatron, installed in the physics building that fall,8 was the first betatron to be installed in any university or hospital in Canada. After calibration, the first patient was treated in March 1949 with the gamma rays emitted (so-called bremsstrahlung radiation) when high-energy electrons interact with a heavy metal target. As indicated above, Johns also wanted to develop 60Co as a more economic and effective source of high-energy gamma rays, and he was able to obtain a 60Co source in 1951 after applying to NRC. The first patient was treated in October 1951; her cancer was cured and she lived to be ninety. This cancer radiation therapy unit continued to operate for more than seventeen years, during which over three hundred patients were successfully treated. This is a very important Canadian example of the use of pure science for very practical medical purposes. Combined with the first 60Co radiation unit in Canada, built by Johns and associates in Saskatoon, research was undertaken in the areas of radiological physics and radiation chemistry on plants and animals. But the main reason for obtaining the betatron was for Katz’s nuclear physics program.9 He developed this program using mainly the bremsstrahlung gamma radiation from the high-energy electrons striking a heavy metal target such as W. The high-energy gamma rays from the W target strike a metal (such as copper

The University of Saskatchewan  11

[Cu]), and a neutron (n) is emitted to give a so-called photonuclear reaction. For example, 63Cu + γ = 62Cu + n.* The high-energy gamma rays in the above example were absorbed by the 63Cu, making it unstable and causing the emission of a neutron. Katz became famous internationally for his studies on the characterization of the resonance energies and cross-sections for the above reactions for a host of different elements. He noticed “breaks” in the neutron emissions as a function of gamma ray energy, and this led to a detailed understanding of the stability of nuclei, and to the well-known nuclear shell model of the nucleus. Writing in 1982, Michael Hayden, the USask historian, recognized even then that talent breeds talent, and that it therefore became easier for USask “to attract outstanding scientists, at least for a while.”10 As we will see in a few pages, Hayden became much less complementary about future big science facilities at USask. By the end of the 1950s, a survey of scientists rated USask in the top ten in North America.1 This should be compared with the surveys done today, in which USask is not in the top ten universities in Canada, not in the top one hundred in North America, and not in the top four hundred worldwide. But, although a relatively small university, USask maintained many outstanding research strengths – for example, in several areas of physics (such as the accelerator program), geology, and agriculture. Its strength in physics was, of course, critical for USask to obtain the huge CLS facility. In the early 1960s, Katz proposed a 140 MeV linac, a linear accelerator to produce 140 MeV electrons (still used by CLS as the injector; see appendix 1), mainly to extend the photonuclear experiments to many other heavier nuclei, and to perform unique electron scattering experiments.11 The cost was $1.75 million, split between NRC ($750,000), USask ($750,000), and the province of Saskatchewan ($250,000). Katz had to use all his great powers of persuasion with university administrators (e.g., President John Spinks) and government politicians (e.g., Premier Tommy Douglas) to get that accelerator and the Saskatchewan Accelerator Laboratory (SAL) funded from multiple sources.† * Note that the same type of reaction is now being used close to seventy years later by the CLS spin-off company Canadian Isotope Innovations Inc. for producing medical isotopes for the world market. † Confirmation of these costs came from two articles. The first source is Ned Powers’s interview with the ninety-year-old Leon Katz, published on January 11, 2000, in the Saskatoon Star ­Phoenix. Katz is quoted as saying “the National Research Council offered $750,000 for the cost of the accelerator,” and he also confirmed that he received $250,000 from the province after a personal meeting with Premier Douglas and President Spinks. The second source, Klines’s ­article,1 reiterates that the province contributed $250,000 and that the total cost was $1.75 million. However, the acknowledgements at the end of the first paper on the SAL f­ acility11 states the following: “we would like to thank the Atomic Energy Control Board for financing the purchase of the accelerator.” This strongly indicates that even the accelerator was at least partly funded from a loan.

12  The Canadian Light Source

Accelerator facilities across Canada funded in the 1950s and ’60s at larger Canadian universities had similar costs, but the contribution of $750,000 from USask was unusually large.* To put that amount in perspective, the total USask operating budget in 1962 was just over $6 million.† The USask contribution was 12.5 per cent of the university’s operating budget, and there was considerable opposition to funding SAL from within USask, as would be expected at any university. As told to me by Glen Caldwell, a member of the USask geology department at that time and later vice-president research at Western, part of the university community demonstrated against the funding of SAL, claiming that USask would lose its library if SAL was funded. The funding (and later success) of the linear accelerator in 1962 has to be one of the most important events in this seventy-year story. Without SAL there would be no CLS at USask today. The most costly Canadian accelerator facility, the Tri-University Meson ­Facility (TRIUMF) operated initially by the three universities of British Columbia, Simon Fraser, and Victoria, was funded in 1968 with $30 million from the federal government. This was the largest and most expensive university facility in Canada until CLS was funded in 1998. TRIUMF is now overseen by twenty Canadian universities. An image from the sod-turning ceremony for the new facility (figure 1) shows Sir John Cockcroft (a British 1958 Nobel Prize in Physics winner awarded for his studies of nuclear transformations using high-energy protons) along with John Spinks, then president of USask. Michael Hayden has claimed that the funding of this facility was the start of decades of decline at USask. Here we see the classic competition and confrontation, common at all universities, between the humanities and sciences for funds and prestige. The traditional statement from the humanities (and from many other disciplines) is always, “If you fund this project, my department and the whole university will suffer from lack of funds.” This theme will appear several times in this book. There is sometimes a lot of truth to the above comment, and the funding of SAL probably had a short-term negative effect on other budgets at USask. But certainly for the much larger CLS, this statement is untrue. New federal and provincial money has funded virtually the whole of CLS with no new capital or operating money from the central university budget over the $1 million per year that was spent on SAL. Moreover, the decades of decline at USask (if that is even true) was partly centred around the relatively poor economy in the 1970s and ’80s and the lack of any population growth that made Saskatchewan and USask relatively poor compared to other provinces * I am grateful to Walter Davidson of NRC for his input on accelerator costs in Canada during this period. † This budget was kindly obtained through the co-operation of Greg Fowler, USask vice-­ president of finance, in 2016.

The University of Saskatchewan  13 Figure 1.  Sod-turning for SAL, May 10, 1962. Nobel Prize winner Sir John Cockcroft from England (left) and John Spinks, president of USask (right).

and Canadian universities. Peter MacKinnon, president of USask from 1999 to 2012, outlines several other significant problems at the university, such as incomplete funding from the province, lack of competitiveness and focus, and poor management.12 SAL became operational in 1964 (figure 2), and was upgraded from 140 MeV to 220 MeV in 1975 and to 300 MeV in 1983 in order to remain competitive with many other newer accelerators across North America and internationally. There is always a continuous struggle (and great competition) to maintain outstanding performance for such a facility in Canada. Upgrades and improvements need to be made continually; and completely new facilities are usually required every twenty or twenty-five years for a facility to maintain funding. The photonuclear studies outlined above were expanded greatly at SAL, and various types of new experiments were begun with new faculty members such

14  The Canadian Light Source Figure 2.  Leon Katz at the new LINAC at SAL, 1964.

as Jack Bergstrom and Dennis Skopik. For example, Jack Bergstrom started unique electron scattering experiments using the electrons from the linear accelerator. The electron beam was targeted onto nuclei such as 15N or 18O, and the electron energies varied, with the scattered electrons recorded as a function of electron energy.13 At a particular electron energy, a “giant resonance” was observed leading to a “shattering” of the nuclei. The giant resonance position increased with an increase in mass, and the nature of nuclear forces could be investigated. The SAL accelerator functioned well and the published work was well known in the nuclear physics community in Canada and internationally. And perhaps most importantly for the present story, the accelerator, detector, cryogenic, and electronics expertise that evolved was crucial for the future development of CLS. Also very important was the project management from Underwood, McLellan and Associates (UMA), initially a Saskatchewan-based engineering company. UMA was critical for the construction of the later accelerators – EROS in the 1980s and CLS beginning in 1999. Leon Katz was awarded many honours; he was appointed a fellow of the Royal Society of Canada and an officer of the Order of Canada. He lived to the

The University of Saskatchewan  15

ripe old age of ninety-four, dying in 2004, and the authors saw him during the CLS construction period, in approximately 2000. He was thrilled, as he should have been, with the development of CLS: his incredible efforts in the 1940s, ’50s, and ’60s had resulted in the impressive CLS facility, much more important scientifically (and more expensive) than the original 1946 betatron. c)  New Electron Accelerator, EROS: 1970s and ’80s Nuclear physics facilities, both in Canada and abroad, were beginning to be defunded in the 1970s; and even at the relatively high electron energy of 300 MeV, SAL was not projected to have a long life without new innovations: no such facility remains competitive for more than twenty years or so. NRC, and then the newly formed (in 1978) Natural Sciences and Engineering Research Council (NSERC), only funded big, truly outstanding international-class projects, and would only fund operating costs if such projects remained internationally outstanding. (Note the very intense competitive pressure for science funding in Canada, and indeed in all countries.) So it was critical that new and more advanced types of accelerators be designed and funded to keep the outstanding group of approximately thirty scientists and technicians together at SAL. So, in the early 1970s, the theoretical group headed by Roger Servranckx, later head of the mathematics department at USask, designed a “pulse stretcher” storage ring (PSR), a semi-circular synchrotron that would dramatically increase the duty cycle of the electron beam and decrease its energy spread for more novel nuclear physics experiments. This new facility was called the Electron Ring of Saskatchewan (or EROS). The first proposal to NRC for this new accelerator (for $4.5 million plus $1 million for a new building) was submitted in 1976, but it was not funded by NSERC until 1983, after the third proposal was submitted. EROS became known locally as SORE, because of the angst caused by numerous failures to acquire funding! In the 1970s and ’80s there were still several electron accelerators in universities across Canada. Many of these were no longer producing internationally competitive research. NRC and NSERC (with the aid of well-known international accelerator physicists) made the very difficult decision to defund most of these accelerators. Indeed, the only electron accelerator in Canada that was still funded by NSERC in 1985 was EROS, which in 1983 had received $5.85 million. NSERC also provided an operating grant of over $2 million per year, provided that the university fund $1 million a year in operating costs. This was used mainly for the power bill. EROS would have been the first of its kind in the world, but the long delay from design in 1971 to construction in 1983 enabled Tohoko University in Japan to complete its 150 MeV PSR in 1981. But SAL had dodged another bullet: without the new funding in 1983, SAL (and most of its approximately thirty staff) would have disappeared very quickly from USask, in which case there would have been no CLS at USask.

16  The Canadian Light Source

The EROS project was initially led by Henry Caplan at SAL, who later ­ ecame chair of the physics department.13–15 Dennis Skopik became the direcb tor in 1987. To save building costs, this ring was ingeniously redesigned in 1982 by Jack Bergstrom, Henry Caplan, Roger Servranckx, and B.E. Norum from the University of Virginia;14 it was then mounted to the basement ceiling of the SAL building, directly above the 1962 LINAC. EROS expanded the nuclear science program greatly because it gave a huge increase in “duty cycle,” making it feasible to perform coincidence experiments. These are experiments in which two or three particles are detected simultaneously (within a few billionths of a second of each other), implying that they all participated in the same reaction with a particular nucleus. Large detectors were constructed to fully utilize the EROS capabilities. One of them was “Igloo,” a neutral pion spectrometer consisting of three tons of lead glass.16 Pions are subatomic particles that are affected by the “strong interaction,” the force that holds nuclei together. The study of pion production after interaction with photons of known energy was world leading at the time, yielded important new information on the strong interaction, and gave the group great expertise in detector and electronics design. EROS also produced synchrotron radiation (as all synchrotrons do with bending magnets to get the electrons to go around in a “ring” structure; see appendix 1). Bill McGowan and I had first talked to Roger Servranckx, the main accelerator designer in Saskatoon in 1973–4, when Bill was proposing the first dedicated 1.2 GeV Canadian synchrotron ring. Indeed, the preliminary design of the new Canadian synchrotron proposed by McGowan came from Roger Servranckx and Ed Rowe, from the Tantalus facility at the University of Wisconsin, combined with McGowan and Hans Froelich in the Western physics department. Servranckx came to London several times in the early 1970s to discuss this design – the first interaction between Western and USask on synchrotrons. But at this time, no one in Canada had yet done any experiments using synchrotron radiation (SR).* This was obviously a high-risk and potentially very expensive project (approximately $10 million at the time), but it was obvious to McGowan and me, even at that time, that this would be incredibly important for Canadian scientists, partly because many countries were already building these facilities to support outstanding novel research. When it became clearer that this was going to be important, the second Western/USask interaction occurred in 1979, when I went to a meeting in Saskatoon (only my second time in Saskatoon since leaving the city shortly after my birth in 1942) to support the second EROS application to NSERC in which the EROS group suggested that

* To put this in perspective, initial experiments with synchrotron radiation began in the late 1960s at several places around the world, including Stanford and the University of Wisconsin in the United States.

The University of Saskatchewan  17 Figure 3.  Les Dallin (left) and Mark Silzer (right), ca. 1990.

the SR from EROS could be a useful radiation source. However, the quality of the synchrotron beams was not competitive, even with sources at Madison that I had already used. Unfortunately, this second application to NSERC for EROS was again turned down. The third application was finally successful in 1983. Other very important players in this story at USask were Les Dallin and Mark Silzer, pictured in figure 3. Les came to USask in the early 1970s and completed his BSc and MSc before doing his PhD in the late 1980s (graduating in 1990). He worked on the EROS design for his PhD with Jack Bergstrom and Roger Servranckx, absorbing all the vast knowledge required to design storage rings of different types.17 Roger Servranckx was winding down his interest in CLS in the mid- to late 1990s, was retired by 1997, and moved to British Columbia. As a result, Les was the principal designer of the CLS electron accelerators, with a lot of checks and balances from accelerator designers from the SAL lab and from all over the world. Mark also had a very important role in synchrotron developments such as the radio frequency (RF) components over the last twenty-five years.

18  The Canadian Light Source Figure 4.  Dennis Skopik outside the front entrance to SAL, ca. 1990.

d)  The Initial Push for CLS, with University Conflict By 1990, it was becoming apparent to the new director of SAL, Dennis Skopik (figure 4), that the program might lose its funding from NSERC – an operating grant of $2.5 million a year, by far the largest research grant at USask. The loss of this grant would again mean the death of the whole nuclear physics program at that school, and the loss of over thirty talented scientists and technicians. Already, the nuclear physics program had escaped two near-death

The University of Saskatchewan  19

experiences – in 1962 and in 1983; and it would be extremely unusual to escape this time. The majority of the group wanted to keep the nuclear physics program if at all possible: science with the photon beams produced from the electrons in a storage ring was very different science than nuclear physics done with electron beams, and scientists rarely change their field of research so dramatically. The nuclear physics program examined the forces in nuclei, whereas the SR experiments studied electronic and atomic structures. But pragmatically, Skopik saw that the only way of keeping the personnel together was to convert SAL to an SR source with completely different science than nuclear physics. To that end, he became a great supporter of the Canadian Institute for Synchrotron Radiation, an organization initiated by Bruce Bigham at the Chalk River Nuclear Laboratories in 1991 (see chapter 4). But it became apparent that former SAL director Henry Caplan (and at the time, chair of physics at USask) would not help Dennis pursue the new project. For example, Caplan refused to give Skopik time off from teaching in the physics department to put in the huge effort to develop the USask proposal. By contrast, I was given a year off teaching to develop the Western proposal. Dennis Skopik was criss-crossing Canada developing support for the project between 1995 and 1999 – a Herculean task that he did remarkably well. The relations between Caplan and Skopik (and other physics faculty members working on the new project) were so bad that Henry was not even invited to the May 1999 party marking SAL’s formal dissolution and incorporation into CLS. Meanwhile, Dennis was also coping with a relatively modest salary (which was not augmented during his monumental work schedule), the exhausting travel to Ottawa and across Canada, the negative messages from the university community, and constant conflict with the provincial representatives who wanted to control and manage the project themselves. Michael Hayden writes that “it is ironic that the campaign to obtain this significant piece of research equipment [CLS] began with the well-intentioned efforts of the head of SAL [Dennis Skopik] to fund jobs for his research staff when NSERC announced that it would no longer fund the accelerator.”1 What is ironic about this? What really competent leader would not try to protect the jobs of thirty-five talented and dedicated colleagues with unique expertise in Canada? Obviously, there was significant financial risk for the university in the CLS project initially; but by the time Hayden wrote that comment, the financial risk to the university was extremely small; and as indicated above, the university has put no new capital or operating funds into CLS in the last fourteen years over what they were giving SAL (approximately $1 million a year). Fortunately, Skopik received very good support from USask president George Ivany, Dennis Johnson, and key Saskatchewan people like Doug Richardson from Mckercher Mckercher and Whitmore LLP, the law firm representing USask at the time (figure 5). It was still true that the odds of getting the new facility at USask were

20  The Canadian Light Source Figure 5.  Doug Richardson (left) and Dennis Johnson (right) at SAL, 1999.

extremely small, but these odds were still far better than the zero chance of maintaining the SAL facility. The number of stars that had to line up in the 1990s for CLS to succeed at USask was incredibly large. When Dennis Johnson and Dennis Skopik spoke about the project at many local, provincial, and federal venues, Skopik would show an overhead slide consisting of a semi-circular probability metre going from 0 to 180 degrees with a moveable arrow on the mid-point of the x axis. In 1992–3, the metre was at about 10 degrees – very low probability. Over the next six years, the arrow slowly moved upward and clockwise; and by the time of the CFI funding in the late 1990s, the metre showed 180 degrees, or 100 per cent probability.

3 The University of Western Ontario: The Beamline and Experimental Expertise, 1970s–1990

a) Introduction to the Canadian Synchrotron Radiation Facility outside Madison, Wisconsin As stated in the introduction to this volume (and shown in appendix 1), a synchrotron facility requires accelerator, beamline, and experimental expertise. The accelerator expertise existed at USask through SAL, while the latter two were developed by several Canadian academic scientists at US facilities. The first and most prominent of these US-based facilities was the Canadian Synchrotron Radiation Facility (CSRF), which began operation with one soft x-ray beamline in 1982 at the low-energy 240 MeV Tantalus synchrotron, built at the University of Wisconsin outside the state capital of Madison. Tantalus was initially designed for nuclear physics and is normally labelled a first-generation synchrotron. Tantalus was converted in the mid-1960s to be the first dedicated SR source (so-called second-generation) in the world. As initially intended, the CSRF beamline was transferred to the “new” second-generation 1 GeV Aladdin synchrotron in January 1986 (see appendix 1) after Aladdin was delayed for over three years due to technical problems. Three soft x-ray CSRF beamlines were funded by large grants to Western between 1979 and 1995. Several hundred Canadian scientists, along with many international colleagues, routinely and successfully used these Canadian facilities beginning in 1983. The two oldest CSRF beamlines were duplicated at the third-generation CLS in Saskatoon, with great improvements in intensity and resolution. The newest CSRF beamline was repatriated to CLS from Aladdin in 2004, and modified for greatly enhanced performance. Several CSRF employees are now employed at CLS; and many students, post-docs, and faculty that did a lot of their research at CSRF are now CLS users. NSERC and NRC funding for CSRF terminated March 31, 2008, when it formally ceased operations. The Aladdin synchrotron and the Synchrotron Radiation Center (SRC) in Madison closed in 2014 (after operating for thirty years) because, as

22  The Canadian Light Source

a second-generation synchrotron source, it was no longer competitive with the new third-generation facilities. The history of CSRF was marked first by excellent co-operation and teamwork from a large number of very talented Canadian faculty, post-docs, and students from a number of universities in Canada; second, by remarkable international collaboration and co-operation from the SRC outside Madison; and third, by a number of good decisions (with a significant amount of luck) on the selection of CSRF beamline equipment (which was not routinely available commercially at the time) and beamline personnel. These three characteristics, along with excellent support from NSERC and NRC after 1988, enabled CSRF to be a very cost-effective, world-class soft x-ray synchrotron facility for over twenty years.4 Of greater importance, a lot of outstanding science resulted at CSRF. b)  My Background in Spectroscopy As a BSc and MSc student at the University of Manitoba in the early 1960s, I was fascinated by existing spectroscopies (such as infrared spectroscopy pioneered by Gerhard Herzberg), which are very useful for characterizing matter; and I was especially interested in the theory of these spectroscopies. Already in the mid-1950s, there were many established spectroscopies, but many others were discovered in the late 1950s and ’60s. Because of the incredibly wide energy range of light in the so-called electromagnetic spectrum that can be used as sources for spectroscopy (from radio frequencies at low energy, through microwaves, infrared, visible, ultraviolet, x-rays, and gamma rays at high energy), dozens of spectroscopies are now known and widely used. In 1963, while looking for a topic and place for my PhD, I stumbled across the first book on a new spectroscopy known as “Mössbauer spectroscopy,” named after Rudolph Mössbauer, the German physicist who discovered it by accident in 1957 and won the Nobel Prize in Physics shortly after. I had not heard of SR, which was discovered in 1947 but was not used until the late 1960s for spectroscopic studies. From that day on, I became very interested in this new Mössbauer spectroscopy, which used gamma rays from radioactive sources such as cobalt (57Co) or tin (119Sn). After a Shell Scholarship enabled me to study at the University of Cambridge, I contacted an old friend, Ian Smith from Winnipeg (who began his PhD studies at Cambridge in 1962), about possible research opportunities at Cambridge. Fortuitously, he found out that a chemistry department radiochemist, Dr. A.G. Maddock, had an opening for a new graduate student. And as luck would have it, he was planning to build a Mössbauer spectrometer. It appeared that I had ended up in the right place at the right time. Dr. Maddock welcomed me to his group in October 1964. In my first week at Cambridge, he scribbled down on the back of an envelope what he considered would be the important components of this new Mössbauer spectrometer,

The University of Western Ontario  23

some of which he had already ordered. Fortunately, with a lot of help from the electronics shop in the chemistry department, in about six months I was able to get one of the first automated spectrometers working well. Since it was a novel technique, several chemists and mineralogists soon became very interested in characterizing iron-containing chemicals and minerals using this spectroscopy. It was the beginning of several very productive collaborations (for example, with Roger Burns, Michael Clark, and Martin Mays); these eventually led to many publications and, in 1973, a book.18 It is exactly this type of research with my collaborative partners that initiated a valuable strategy that I still follow today: collaborative research in several different scientific fields, with both very basic components (including simple theory) and very applied components. For example, with any new spectroscopy there are opportunities to develop the spectroscopic method – by discovering new phenomena and devising simple theories to explain and predict both the new features (peaks) and the previously known features. A good understanding of spectroscopy can then be applied to gain an enhanced understanding in both basic and applied research areas. I have always felt strongly that it is possible to do outstanding basic and applied research, and that one enhances the other. c) Bill McGowan and I Begin to “See the Light,” 1970–8: The Centre for Chemical Physics I arrived in the chemistry department at Western in July 1970, where I continued a strong research program using Mössbauer spectroscopy18 on solid state iron and tin compounds and minerals. But, having already realized that Mössbauer spectroscopy was severely limited to solid state studies on just a few elements such as iron and tin, I wanted to get involved with a new spectroscopy that could be used to characterize all elements in bulk solids, surfaces, and gases. Before leaving Cambridge in 1970, I had attended an Inorganic Chemistry Faraday Society Discussion there and presented a paper on the use of Mössbauer spectra for obtaining bonding and structure information in iron (Fe) compounds. This meeting also featured early papers on the use of photoelectron spectroscopy (using both x-ray and ultraviolet sources) for studying bonding and structure in inorganic compounds in the gas and solid phases. I became very interested in this new spectroscopic area;19,20 and after setting up a Mössbauer program at Western in 1970–1, I began a plan to purchase the second commercial photoelectron spectrometer, a McPherson ESCA 36 (ESCA standing for Electron Spectroscopy for Chemical Analyses) to perform both UPS (Ultra-Violet Photoelectron Spectroscopy) and XPS (X-ray Photoelectron Spectroscopy) on gases and solids. My progress toward getting an instrument at Western was greatly enhanced by a fortuitous meeting with John Thomas, a visiting professor from Aberystwyth, Wales, in early 1971. Professor Thomas had just bought the first

24  The Canadian Light Source

commercial XPS instrument. I gave him some mineral samples from a previous collaboration with Roger Burns18 (who was now a professor at MIT) and a joint paper was published. It was not groundbreaking, but it led to some excellent work with Wayne Nesbitt at Western almost forty years later with a new, more sophisticated instrument. This spectroscopy had been developed in the 1960s, mainly by the Turner group at Oxford University19 in England and the Siegbahn group at Uppsala University in Sweden.20 Kai ­Siegbahn won the Nobel Prize in Chemistry in 1981 for the development of XPS. In early 1972, I mentioned to my department head, Howard Clark, that I would like to apply to NRC for a photoelectron spectrometer at Western. I was still not aware of SR. Clark strongly supported a $140,000 application (a very large NRC grant at the time, but much smaller than the SAL LINAC funding in 1962 at USask) to NRC in October 1972. He sent me to a photoelectron workshop in Cardiff, Wales in the summer of 1972, and also did a substantial amount of the grant writing. He suggested that the operating funds (including a technician’s salary) for the facility could come from industrial business, and he also suggested that we get support from other interested universities to strengthen the application. It is important to emphasize that Clark was not directly interested in the technique, and he was not himself on the application. Such support from a senior administrator is not common; but Howard Clark was not a “common” leader, and he generously helped many of the new chemistry faculty reach their full potential with his time and helpful suggestions. I received support from the Universities of Windsor and Toronto for the South Western Ontario ESCA facility, the grant was awarded in March 1973, and the McPherson ESCA 36 spectrometer was working very long hours with a full-time technician, Bob Lazier, by early 1974. This same versatile spectrometer was purchased by other groups in Canada (for example, Ron Cavell at the University of Alberta and Stewart McIntyre at Atomic Energy of Canada Ltd., in Pinawa, Manitoba) at about the same time. The Western instrument supported very diverse pure and applied science programs using ultraviolet and x-ray photons on gases, solids, and surfaces, including industrial research with support from steel companies such as DOFASCO, which paid for Bob Lazier’s salary. This mixed academic-industrial approach was very unusual at the time. Fortuitously, Bill McGowan, the first director of the Centre for Chemical Physics (CCP) at Western, initiated the Canadian SR effort just as I began planning for the photoelectron purchase in 1972. Gerhard Herzberg and one of his close NRC spectroscopic colleagues, Alec Douglas, along with the director of the Tantalus SR facility in Madison, Ed Rowe, attended the first Canadian workshop organized by Bill McGowan on the uses of SR at Western in October 1972. At this workshop, I realized the power of SR to improve dramatically the performance of XPS compared to laboratory sources; and more importantly, that SR could be useful to a wide range of disciplines in science,

The University of Western Ontario  25

medicine, and engineering.21,22 McGowan, with his tremendous dedication and aggressive energy, pushed extremely hard for a Canadian synchrotron between 1972 and 1977. For example, he applied in July 1973 to NRC for a feasibility study for a Canadian SR source. But there were no known Canadian users at the time; and synchrotron research had only begun about 1968 at Wisconsin, Cornell, Hamburg, and Tokyo. This and later requests were rejected. In January 1975, McGowan hosted a CCP SR workshop to consider a design from Roger Servranckx of USask for a 1.2 GeV Canadian low-energy synchrotron (similar to the one being designed at the University of Wisconsin). McGowan suggested to the university’s senior administration a possible site for the synchrotron at Western. It is important to comment here on CCP. It was formed in 1970 by a large grant from NRC to Bill McGowan and Pat Jacobs from the Western physics and chemistry departments, respectively. This centre was critical for the development of not only Canadian synchrotron studies, but other large Western groups/centres in surface and materials science, such as Surface S­ cience Western (SSW) in 1981 and Interface Science Western (ISW) in 1986. CCP was one of the first academic science centres in Canada to emphasize interdisciplinary and industrial research, and its members consequently came from physics, chemistry, applied mathematics, engineering, and several medical areas. It hosted many interdisciplinary lectures and workshops; but the most important initial impact came from the international fellowship program that paid several senior international faculty to spend their sabbatical years at Western. Several of these international visitors were very important contributors to the excellent research done at Western between 1970 and 2000, as well as the very large research funds associated with those efforts. d)  The First Canadian Users of SR Canadian scientists began to use synchrotron radiation in the United States beginning in 1974. The first was Daryl Crozier from Simon Fraser University. He established a beamline partnership at the Stanford synchrotron in 1974 for high-energy extended x-ray absorption fine structure (EXAFS) studies.23 In late 1974, I made an application for beamtime to Ed Rowe and the Tantalus synchrotron at the SRC at the University of Wisconsin Physical Sciences Laboratory to obtain photoelectron spectra of volatile organometallics on Tantalus, the first dedicated second-generation synchrotron source in the world.24 For one week in March 1975, I went to Madison with my new PhD graduate, T.K. Sham, to collaborate with Dean Eastman from IBM, who was very well known for recent synchrotron-based photoemission studies of semiconductors. Indeed, there was a large group of researchers from both academia and industry

26  The Canadian Light Source

that were doing outstanding semiconductor research at Tantalus, including Gerry Lapeyre from the University of Montana. Eastman had a ultra-high vacuum (UHV) photoemission system, and was extremely flexible and helpful in obtaining spectra of organometallic films by subliming very thin films of T.K. Sham’s volatile compounds (from his PhD thesis) on metal substrates to avoid charging problems. These compounds were very incompatible with the UHV system, and most researchers would not have allowed these compounds to be anywhere near their equipment. This one week of very intense work changed my career. I decided to do part of my sabbatical (July 1975–January 1976) working with Eastman (who was later made senior vice-president at IBM) and post-doc Wolfgang Gudat (later director of the Berlin synchrotron); and I was greatly encouraged by Eastman and Rowe to spend at least six months at the Tantalus facility. Those six months transformed my scientific career, and indeed the Canadian synchrotron effort. I began my year’s sabbatical (six months in Madison and six months in Uppsala, Sweden) in the summer of 1975 with my wife, Joan, and two young children, David and Catherine. We rented an apartment near Lake Monona, about a twenty-five-minute drive from the Tantalus facility south of Madison. Fortunately, my wife was incredibly flexible and co-operative and looked after our young family full-time in the most unselfish way. The positive international co-operation at the lab was very noticeable, mostly because of Ed Rowe. He welcomed Canadian users heartily, partly because he had Canadian relatives. Undoubtedly, he hoped that a strong Canadian group would enhance the research output of his facility and give him better possibilities for an increase in funding from the National Science Foundation for an underfunded facility. The senior staff, including Rowe and Charlie Pruett, did not have PhDs, and were not properly acknowledged at the University of Wisconsin for their fine work in getting Tantalus operating and their efforts to encourage users. The management, my collaborators Dean Eastman and Wolfgang Gudat, my outstanding graduate student and post-doc T.K. Sham, and the technical staff were very approachable and helpful. There was also a good deal of enthusiastic humour, often involving Ed Rowe. Exciting new results were often obtained, but only after waiting at the synchrotron console for a few hours while the technical staff tried desperately to get the electrons to store in the storage ring. In contrast, at modern facilities such as CLS, the electrons are stored in seconds. Operation of a synchrotron was quite an art at the time, and the Tantalus ring was funded at a very poor level, making improvements very difficult. So great patience was often required. In addition, other users had extremely varied scientific interests; as a result, I learned a lot about related fields from nearby users working on other beamlines. This cross-fertilization of ideas from researchers in different disciplines is also critical for outstanding research. Most synchrotrons worked twenty-four hours a day (as does CLS today), but Tantalus was working only from 9 a.m. to

The University of Western Ontario  27

6 p.m. because of lack of funding. Long days were still usual because samples had to be prepared in the evenings. We were able to do some good science in that six months.25 In addition, I learned a lot about synchrotron beamlines and had good discussions with the beamline and optics groups at Tantalus, especially Charlie Pruett. These discussions were important for the future applications for Canadian beamlines. e)  Personal Stories: Humour and Other Activities As mentioned above, Ed Rowe had a great sense of humour, and this was very important to the overall positive atmosphere at SRC. One manifestation of this humour showed in his initiation and presentation of the unique G.J. Lapeyre Award for the scientist who had “met and overcome the greatest scientific or technical obstacle in a given year while working at the SRC.”24 Gerry Lapeyre, from the University of Montana, was an outstanding solid state physicist who did some classic spectroscopy on semiconductors at the SRC in the early 1970s. In 1972 or 1973, the underground vault that held the synchrotron was being enlarged and the temporary roof was inadequate to hold back significant rain. During a facility picnic, a severe thunderstorm resulted in the flooding of the experimental vault. Gerry had important equipment inside, and he rushed back from the picnic to dry out all the electrical connections to his and his colleagues’ equipment. Because he had a farming background, he also used the lab’s tractor to build earthworks to divert the water. Gerry then had to deal with the firemen and county deputy sheriffs who were notified of electrical problems. Ed Rowe created the eponymous award to commemorate Gerry’s great service to the lab. Gerry, of course, received the first award, which consisted of a Leinenkugel beer can mounted on an octagonal wooden base with a piece of concrete from the vault sticking out the top of the can. This unique trophy was presented by Ed Rowe at every SRC annual meeting for the next nineteen years; and his extremely humorous introduction of the award (which went on for over five minutes) was the highlight of every meeting. After 1978, Ed focused a lot on perhaps the most deserving of the award winners, the 1978 recipient Mort Traum, from Bell Labs. Traum had been “attacked” by a large welding clamp, which attached itself to his very sensitive manhood! Fortunately, he was not injured badly and was quickly discharged from the hospital. In a very humorous act, Ed speculated on how this happened; indeed, the story was exaggerated at every subsequent annual meeting. Our Canadian contingent was always active at these very enjoyable gatherings. And in fact, our Canadian employee at Tantalus, Kim Tan, won the award in 1983 for his important efforts to move and align the Canadian beamline at least twice in the previous year. The time in Madison was enjoyable in other ways as well. I curled competitively (as I had done since the age of eleven) at the very old Madison Curling

28  The Canadian Light Source

Club; Joan and I also sang in the large Madison-area community chorus and attended a local Methodist church. We drove back to Winnipeg in mid-January before going to Uppsala on the second leg of the sabbatical. I was fortunate to take a curling team, which included my father and father-in-law, into the very large Winnipeg bonspiel in mid-January, with over six hundred teams competing. We played eight games in three days at curling clubs across Winnipeg between 8 a.m. and midnight. The highlight was beating a young team skipped by Mark Olsen, who later played on Brier-winning teams with Kerry Burtnyk. f) Six Months in Uppsala, the First Proposal for a Canadian Synchrotron at Western, and the Proposal for a Beamline at CSRF in Madison The time in Uppsala in Kai Siegbahn’s lab20 was even more interesting, in many ways because of the cultural differences. This was, after all, our first experience living outside of the English-speaking world. My Swedish colleagues (such as the graduate student Svente Svensson that I worked with directly) were very helpful and collegial, but usually there was less high-spirited humour than there had been in Madison. Joan and I went to Swedish language classes twice a week for about two months. I then started to travel to other European cities, such as Namur, Belgium and Hamburg, Germany, where I visited Wolfgang Gudat to finalize papers from Madison. The science went well,26 although not as well as hoped; but the experience gave me a lot of ideas for synchrotron research, which I would put to use when CSRF in Madison began working three years later. In September 1975, Bill McGowan submitted a very large proposal, entitled “A National Laboratory for SR Studies,” to Paul Redhead, the director of the Physics Division at NRC in Ottawa. In response to this proposal, Redhead chaired an ad-hoc NRC committee on SR; and that committee met in June, August, and October of 1976. I returned to Ottawa from Sweden for the first two meetings. The Canadian chemists on that committee were Camille Sandorfy (University of Montreal), Chris Brion (University of British Columbia), and me. Bill McGowan also organized the first very large international SR workshop at Laval University in June 1976. There were fifty-five attendees at this workshop, though just eleven Canadian scientists. A 624-page report on the many uses of SR was submitted to Paul Redhead’s ad-hoc committee. There was still very little Canadian involvement at this time, but it was obvious even then to nearly all the committee members that SR research was going to be very important for Canadian science, and that Canada should therefore be a major player in SR research along with all the other G7 countries that already had SR facilities. However, because there were still very few Canadian users, and because Canada was (and still is) averse to taking “risks,” this proposal was turned down yet again.

The University of Western Ontario  29

When I succeeded Bill McGowan as the next director of the CCP in 1978 (a position I held until 1981), I spearheaded the first application for a Canadian beamline and accompanying equipment (of which more below) at CSRF in Madison. Also, under my leadership, the centre formed SSW (page 25) with several million dollars of equipment from my many grant applications in 1980. The dean of science, John Bancroft (no relation, although we were friendly enough to be relatives), and the president, George Connell, were critical to the funding of both CSRF and SSW, and John Bancroft put in an outstanding effort for a project that was not in his research area. But they both recognized research excellence, and were willing to contribute over $1 million to the development of the two facilities. Bancroft and I hired Stewart McIntyre from Atomic Energy of Canada Ltd. (AECL) in Pinawa as director in 1981 to lead the SSW facility, which performed both academic and industrial research, again perhaps a first in Canada. This facility is still a great self-sustaining success, due largely to Stewart’s talent and dedicated efforts until 2006, and the strong leadership of Leo Lau and Dave Shoesmith since then. The concentration of talent at Western in the surface/materials area increased in 1986 when four outstanding researchers from the Chalk River Nuclear Laboratories were hired (Ian Mitchell, Peter Norton, Willy Lennard, and Keith Griffiths); they brought along about $6 million of sophisticated surface and material science equipment from their laboratories. This also required over $1 million from Western, which was used to set up their laboratories at the school. This investment was resented in some parts of the university, of course, but it led to thirty years of outstanding work, which is still continuing after their retirement. Again, John Bancroft had to fight very hard to get the university to agree to put up another $1 million investment, which was comparable in size to the 1962 USask investment for SAL. Along with the hiring in 1987 of T.K. Sham and Masoud Kasrai for research at CSRF in Madison, Western was in a terrific position to host CLS a decade later. But in 1977 it was still apparent that a Canadian synchrotron was many years away; the next best thing, then, was to construct a beamline or beamlines at a US facility. As the new director of the CCP at Western, I led a $541,000 application (co-applicants were Bill McGowan and Bill Ware, both from Western) to NRC in October 1977 for CSRF to construct one Canadian soft x-ray beamline and three experimental chambers for the existing Tantalus and newly designed Aladdin SR sources at the University of Wisconsin-Madison. I had a very cordial invitation from Ed Rowe (including exclusive use of an SR port on Aladdin at no charge in perpetuity), great technical support at Western from post-docs Raj Gupta and Leighton Coatsworth, and excellent scientific and technical support from Ed Rowe and Charlie Pruett from Tantalus. In July 1978, the newly created NSERC (which took over from NRC in May 1978 as the funder for university research) awarded $320,000 capital for CSRF in Madison: this would cover the soft x-ray beamline ($200,000), and two

30  The Canadian Light Source

experimental chambers (my photoelectron spectroscopy and Bill McGowan’s x-ray microscopy of cells). Note that the cost of this equipment was an order of magnitude less than both the nuclear physics equipment at USask (described above) and the present cost of beamlines at CLS. In early 1978, the NRC committee, chaired by the outstanding laser physicist Boris Stoicheff from the University of Toronto (and the author of the Herzberg biography cited earlier5), complimented the Western group for their initiative “on behalf of researchers in Canada,” and went on to say that “their persistent efforts during the last several years to stimulate interest in and acquire funding for a Canadian Synchrotron Facility have led to the current unique opportunity for Canadian science.” Little did Stoicheff know the magnitude of the opportunity.* As is usual in Canada, not enough capital money was awarded to do the job; it was therefore necessary to arrange multiple sources of funding. This required a lot of extra time and effort with no thought of extra compensation. Bill Ware decided to opt out because of the funding shortfall; Bill McGown, for his part, wanted more money from the start for his experiment. I had to tell him that he had to live with his allotment because the whole project was underfunded. These incidents demonstrate the large problems faced by collaborative projects, which are nearly always underfunded in Canada. As director of the CCP at Western, I was able to obtain an additional $100,000 from the school’s Academic Development Fund with strong support from Dean Bancroft. George Connell, the president of Western, had asked twelve strong research groups at the school, including the CCP, for applications of approximately $1 million each in 1980; these were to be funded from the university’s first big fundraising drive (the so-called Second Century Fund), which marked the hundredth anniversary of Western in 1977. The late starting date for this fund was mostly because Connell only became president in 1979. My application on behalf of the CCP was mainly for the formation of SSW. I included $100,000 in this request to help out with the budget shortfall for CSRF. Another $40,000 came from yet another application to NSERC the next year to complete the two experimental chambers. But the question remained: Who would pay the operating costs for CSRF? As stated in NSERC President Gordon McNabb’s letter of July 7, 1978, NRC was to own, staff, manage, and maintain the facility. This sounded very promising. But then, NRC allotted only $5,000 annually for the ownership, operation, and maintenance of the equipment! After months of correspondence with both NRC and the newly formed NSERC, the $5,000 was not increased. (It is important to note here that both NRC and NSERC became very supportive of CSRF * Interestingly, Boris Stoicheff ’s son Peter, an English professor, became president of USask in 2015. As well, Peter’s wife, Kathryn Warden, was the director of communications at USask, and very capably looked after the communications portfolio at CLS from 1999 to 2002.

The University of Western Ontario  31

financially and scientifically, and NRC personnel such as Norman Sherman, and later Arthur Carty and Walter Davidson, were enormously helpful as well.) This small operating contribution created even more enormous problems for me and for the development of CSRF. Fortunately, Western (with Dean Bancroft, President Connell, and Associate Vice-President Howard Baldwin being enormously helpful) scraped together enough funding by 1980, mostly through the newly announced Second Century Fund mentioned above, mainly to hire Kim Tan in January 1980 for most of five years as CSRF operations manager in Madison. Kim Tan did his PhD with Bill McKonkey at Windsor, and a post-doc with Chris Brion at the University of British Columbia in electron energy loss spectroscopy. Though he had no synchrotron experience he did have a lot of spectroscopic experience using high vacuum conditions, and more importantly a very good attitude. He was the only CSRF employee from 1980 until 1987, and he was the glue that held things together, especially in the first ten years. Without the contribution from Western, the facility could not have gone ahead. These two very large projects (SSW and CSRF), combined with three lengthy and exhausting sessions on the Western tenure committee fighting lawyers in university tenure appeals hearings late into the night, completely exhausted me. In September 1981, I collapsed, and subsequently spent close to six months recovering – fortunately successfully – with the help of a very supportive family and several friends and colleagues (for example, Tom Feasby and Jack Lorimer). During these months, I played the piano for many hours (an important activity for most of my life), and I also took many organ lessons to try to learn the organ pedals properly. It should also be noted that the CCP was set up for collaborative research, but in the CCP’s entire forty-year history, from 1970 to 2010, no other significant collaborative project was launched by the centre’s many directors, other than the two during my three-year tenure from 1978 to 1981. It requires a lot of effort to get these projects going and to successfully apply for multiple research grants for fellow researchers to use. How many people are going to put in that kind of effort? Howard Clark and John Bancroft did at Western, Dennis Skopik did in Saskatoon, and Adam Hitchcock, T.K. Sham, and Daryl Crozier put in great efforts on behalf of the Canadian synchrotron community more broadly. Fortunately, Dennis and I worked very well together, even if we were still competing against each other. All the initial purchases for CSRF were funnelled through NRC, with Norman Sherman guiding them through the NRC purchasing procedures while in constant contact with me. We had a strong and resonant relationship, which is so important for an endeavour like this. The main piece of equipment for any beamline, the monochromator, had two competing bids. The lowest bid came from McPherson Engineering, a well-established company in Boston that I had ordered large laboratory equipment from earlier, and from which I had

32  The Canadian Light Source

a new electron analyzer on order. However, McPherson had not produced a working monochromator. A higher, but still competitive, bid came from a very new company, Baker Engineering in Evansville, Wisconsin (near Madison), that was marketing a unique soft x-ray monochromator (a so-called Grasshopper monochromator) designed and built by Neil Lien and Charlie Pruett of the Physical Sciences Laboratory (PSL) at the University of Wisconsin and Fred Brown, a physicist from the University of Illinois. Our monochromator was the fourth one built (thus the Mark IV label), but the first one manufactured commercially by Baker Engineering. Because I knew the Baker group had produced three successful monochromators, and I knew and trusted the three designing scientists mentioned above, I had to intervene in the NRC bidding process, which would have chosen the lowest bidder, McPherson. But the choice of Baker led to substantial problems with McPherson. Because I had already ordered the electron analyzer for the photoelectron spectrometer from McPherson, the rejection of their bid for the monochromator led immediately to an indefinite delay in the delivery time for the electron analyzer. Needless to say, lawyers became involved in the successful effort to cancel the order for the electron analyzer from McPherson. This monochromator was state of the art in the soft x-ray energy region from 20 eV to 200 eV photon energies, it gave very good spectral resolution, and it had proven to be easy to use, rugged, and dependable at the Tantalus ring.27 This $200,000 instrument worked well immediately, and led to the majority of the publications at CSRF. It was still working routinely in 2007, when the new equivalent opened at CLS. The competing bidder never did produce a working monochromator. My insistence that Baker get the contract over a lower bidder paid off big time. Kim Tan worked closely with Baker Engineering in Evansville. He had excellent co-operation and collaboration with the three monochromator designers mentioned above to assemble the soft x-ray Grasshopper monochromator, and in one year the monochromator was successfully tested on Tantalus (figure 6). The first publication appeared in 198227 and the beamline plus the XPS ordered from Leybold in Germany was working routinely on Tantalus by January 1983. However, we needed the new synchrotron, Aladdin, to get the best results. Aladdin was seriously delayed so that my first, and top-notch, synchrotron graduate student Brian Yates (now at CLS) had to use Tantalus for all his novel gas phase PhD results.28 While Doug Shinozaki and Bill McGowan also had a very strong program on soft x-ray microscopy of biological specimens on Tantalus,29 McGowan soon lost interest. Shinozaki, however, continued the strong research program on soft x-ray microscopy for many years. Of the three initial scientists on the first application, I was the only one who followed through, a not uncommon scenario in academia.

The University of Western Ontario  33 Figure 6.  The official opening ceremony of the first CSRF beamline (the Grazing incidence “Grasshopper” beamline) at the Tantalus synchrotron in early 1983. From left to right: Norman Sherman (NRC, CSRF manager), Bill McGowan (professor, Western), Brian Yates (graduate student, Western), Ed Rowe (Tantalus director, Physical Sciences Laboratory), Brenda Addison (graduate student, Western), Mike Bancroft (CSRF scientific director, Western), and Kim Tan (CSRF operations manager, Western).

Fortunately, the Finnish physicists Seppo and Helena Aksela, sponsored in 1984 by the CCP at Western, wanted to begin synchrotron studies of the rare gas atoms, and our beamline was as good as any in the world at the time for these studies. Seppo doggedly recorded most of the initial resonance Auger spectra of atoms at CSRF on Tantalus, and these spectra were interpreted by Helena. The resulting series of papers on neon, argon, krypton, and xenon are widely cited30 and formed the basis for the Akselas’ higher-resolution studies in the 1990s on higher-intensity synchrotron beamlines in Lund, Sweden. With seven publications in 1984–5 and sixteen publications in 1986, the CSRF facility was “taking off,” and operating money from NSERC and NRC increased greatly. NSERC introduced infrastructure grants in 1983, and Norman Sherman strongly supported increased funding from NRC. The NSERC and NRC grants of $32,000 and $16,000, respectively, allotted in April 1984 came closer

34  The Canadian Light Source

to paying Kim Tan’s salary plus maintenance; and by 1987–8, NSERC and NRC were contributing $82,000 and $35,000, respectively, and Western funding was no longer required. As was the case for SAL at USask in 1962 (see chapter 2, section b), it is usually not possible in Canada to get this kind of ambitious program off the ground without critical university contributions. We were paying nothing for the use of Aladdin; and Kim Tan was very skillful in accessing technical help from the Aladdin facility for very reasonable fees. Indeed, as noted previously, he won the G.J. Lapeyre award for some of his outstanding efforts in 1983.The huge effort required to gain reasonable operating funding again illustrates the difficulties in getting anything really novel started in Canada – or outside Canada in this case. Exchange rate issues (the value of the US dollar varied from CA$1.10 to CA$1.60 over the years) made budgets very difficult, and long-term visa and health-care issues for employees were always problematic. Fortunately, Norman Sherman at NRC and the Western Dean’s Office were always helpful. Although it took between four and eight hours to get to Madison, I always enjoyed about four trips a year for the exciting science with dedicated and talented students and post-docs, combined with rather few problems at Aladdin and CSRF. I especially enjoyed the rural farm atmosphere (cows, tobacco, and corn) surrounding the SRC, the simple SRC residence (owned by the local farmer Mr. Green) 18 kilometres from Madison near Stoughton, and the easy 1.5 kilometre country walk to and from the SRC. Aladdin was supposed to be working in 1982, but multiple problems seriously delayed the commissioning of this new second-generation 1 GeV synchrotron. For example, the concrete floor was poured in four sections, and the sections moved relative to each other, throwing out the alignment of the different parts of the ring structure. Moreover, there was considerable concern by 1985 that the funding for Aladdin would not be renewed, and it was obvious that the Canadian program could not reach its potential on Tantalus. Fortunately, Brian Yates was able to complete his PhD on Tantalus.4,28 By late 1985, most of Aladdin’s problems were solved, and reasonable electron currents could be stored with good lifetimes (see appendix 1 for a schematic of Aladdin). We decided to transfer the CSRF beamline to Aladdin in early 1986; this was one of the first beamlines installed on Aladdin. Ed Rowe still promised us exclusive use in perpetuity. This switch to Aladdin opened up a lot of new science for CSRF users on the Grasshopper beamline, because the Aladdin x-ray intensity was much higher than on Tantalus over a broader energy range from 10 eV to 4 keV. The Grasshopper beamline became useful to over 300 eV photon energies on Aladdin; but most importantly, the higher energies were available using other higher-energy monochromators. Many new users became involved on the Grasshopper beamline, and the scientific program expanded greatly, from the two initial programs mentioned above to absorption studies using XANES

The University of Western Ontario  35

(x-ray absorption near edge structure) of gases and solids of all kinds, along with high-resolution XPS studies of gases and solids.31 Ron Cavell from the University of Alberta generously gave CSRF one of his XPS instruments, which was extensively modified by my dedicated and talented graduate students John Bozek and Jeff Cutler for the highest-resolution molecular gas phase XPS and Auger studies on a second-generation synchrotron. Tolek Tylisczczak (courtesy of Adam Hitchcock) and John Bozek did most of the hardware and software development to scan photon energies using the Grasshopper monochromator for XANES studies. g)  Two New Beamlines at CSRF With the higher energies available from Aladdin, I planned in 1987 (after becoming chair of chemistry at Western in 1985) for a new beamline to cover the energy range from 1.5 keV to 4 keV. Fortuitously again, the Ontario Centre for Materials Research was formed in 1987, and I submitted a $2.1 million application to the centre for this beamline (total cost about $500,000), as well as experimental chambers for solid state spectroscopy and two five-year appointments – a new professor in my chemistry department and a post-doc for the beamline. T.K. Sham, my excellent former PhD student, was hired (after eleven years at Brookhaven National Lab doing synchrotron research) to design the beamline at CSRF in 1987 in expectation of the grant. After comprehensive reviews, the grant was successful, T.K. Sham was hired as an associate professor in chemistry at Western, and Bing Xing Yang from Brookhaven was hired as a post-doc to construct the beamline on Aladdin in 1988. The $210,000 monochromator was built by the University of Wisconsin PSL, headed by Fred Middleton. Bing Xing did a remarkable job getting the whole beamline working in less than eighteen months32 with some help in 1989 from the PSL group, Taiwanese post-doc Jin-Ming Chen, and Swedish post-doc Bengt Olsson. It is interesting to note that Baker Engineering also bid on the monochromator, but their design at Brookhaven National Lab never worked well. In contrast, PSL’s unique double chrystal monochromator (DCM) worked extremely well for almost twenty years. Another hire to my group in 1987 also enhanced greatly the research at CSRF. Masoud Kasrai, an Iranian colleague of mine at the University of Cambridge between 1964 and 1968, became a senior chemical physics fellow at Western in early 1987; he spent over twenty-five years at Western, researching at CSRF for many weeks every year and helping unselfishly in many ways with the many students and post-docs in my and other Western groups. Indeed, Kasrai and T.K. Sham – the two dedicated SR appointments at Western – became absolutely critical to the ensuing success of CSRF, and a lot of excellent pure and applied research resulted. In these days of Trumpian animosity toward immigrants, it is critical to note that T.K. Sham (from Hong Kong), Masoud Kasrai (from Iran), and of

36  The Canadian Light Source

course Bing Xing Yang (China), Jin-Ming Chen (Taiwan), and Bengt Olsson (Sweden), were all from different countries and cultures. Without such international collaboration, science cannot flourish. And without these individual scientists, and the incredible generosity of Ed Rowe in Madison, it is quite likely that CLS would have been delayed several years. Every year, I sent “request for beamtime” letters out to users, as is done at all synchrotron facilities. After consultation with the users, I set up a detailed schedule, with each user group receiving between one and four weeks free beamtime a year on each beamline. Nearly everyone stayed at the residence about 1.5 kilometres from Aladdin. CSRF’s productivity increased enormously after 1990. For example, over forty scientists from more than ten Canadian laboratories and universities did experiments there from 1990 to 1992, and the productivity of the facility increased dramatically to over thirty papers per year in 1992 in a wide variety of scientific areas (see reference 4, table 4). Canadians were guaranteed 80 per cent of the beamtime on the DCM. Users came from the following Canadian organizations: the Universities of Alberta, British Columbia, Carleton, Laval, McMaster, Simon Fraser, Toronto, Waterloo, and Western Ontario; along with the industry and government organizations AECL, Alcan, Canmet, Esso, INRS, and NRC. Foreign users came from the cities of Oulu and Turku, in Finland, the SRC in Madison, and Brookhaven National Laboratory. The research interests of this group varied widely: development of new techniques such as soft x-ray small angle x-ray scattering (SAXS), extended x-ray absorption fine structure (EXAFS) in liquids from photoconductivity measurements, x-ray excited optical luminescence (XEOL), and soft x-ray fluorescence detection; radiation damage of polymers, reflectivity of multi-layer mirrors, XANES/EXAFS studies of a huge range of gases, solids, and surfaces, including semiconductors, coals, glasses, multi-layers, metallo-proteins, and tribofilms; and high-resolution XPS/Auger measurements of simple molecules and organometallics.31 Perhaps the most outstanding research on protective tribological films in car engines was supervised directly by Masoud Kasrai. Many graduate students and post-docs (see theses by Zhangfeng Yin, Marina Fuller, Morey Najman, and Mark Nicholls in reference 4, table 1), with collaboration from Esso Canada and Chevron US, determined the detailed chemistry and spatial characteristics at the micron level of the protective films in car engines that have dramatically increased car engine lifetimes in the last fifty years. This research was necessary to further modify and improve these films. Of the nearly fifty papers that resulted from this research, many are among the most cited (over 200 in two cases, with a total of over 2,500 citations) in the entire tribology literature. The funding for the facility from NSERC and NRC ($173,000 and $50,000, respectively) was now stable and adequate; and as is the international norm, there were no user charges of any kind for use of the two beamlines.

The University of Western Ontario  37 Figure 7.  The official opening of the third CSRF beamline (the SGM beamline) in 1999. From left to right: Ron Cavell, Brian Yates (CSRF beamline manager), Nils Petersen (vice-president research, Western), Walter Davidson (NRC, CSRF manager), T.K. Sham, Mike Bancroft, Kim Tan, and Jim Taylor (the new director of Aladdin). Note that Nils Petersen and Walter Davidson later became chairs of the CLS Board of Directors.

The intermediate x-ray energy range from 300 eV to 1,500 eV was still lacking at CSRF. I applied to NSERC for $1.4 million in 1992 for a third beamline using a so-called spherical grating monochromator (SGM) to give CSRF high-resolution capabilities between 250 eV and 700 eV. This funding was obtained in 1994 after NSERC organized the usual comprehensive review meeting with an international panel, which rated the proposal very highly. My first SR graduate student, Brian Yates, was hired to construct the beamline, and the entire beamline was ordered from McPherson Engineering in 1995.* Brian did a marvellous job, mostly on his own, because there was so little money – as usual – for additional help (see figure 7 above). He managed to hold up to the great stress of getting the beamline working. And though it was somewhat delayed by * This is the same McPherson that we had legal problems with in 1980.

38  The Canadian Light Source

McPherson, it met the specifications in 1998. Almost immediately, many general users began using it.33 This beamline enabled very high-resolution studies at the K edges of carbon, nitrogen, and oxygen. Adam Hitchcock donated his photo-ionization chamber for coincidence measurements, and this was used by Adam, his students, and John Neville from the University of New Brunswick. The Grasshopper and DCM beamlines were still working well at CSRF from 1999 until 2006 under the direction of T.K. Sham and Walter Davidson of NRC. Some of the most influential scientists at CSRF (Adam Hitchcock, Masoud Kasrai, and T.K. Sham) can be seen in figure 8 (next page). In addition, figure 8 also shows Daryl Crozier, the first Canadian to use SR and someone who has been very important in the Canadian SR effort ever since. The SGM was shipped to CLS in 2004, where it formed the basis for a very fine thirdgeneration beamline led by T.K. Sham. It has performed as well as any in the world. The Grasshopper beamline was shipped to CLS as a museum piece, and the DCM line remained at the SRC for general use until the SRC closed in 2014. The DCM replacement at CLS, with T.K. Sham again as the beam team leader, was ready for use at CLS by 2009. After the usual difficult “teething period” from 1980 to 1984 described above, CSRF became a very successful SR facility in the soft x-ray region with a small fraction of the funds required by most international SR facilities. Several hundred researchers used the facility over the next twenty-five years, over 50 PhD and MSc theses were produced (see reference 4, table 1), and over 500 papers were published.* The total capital cost for the three beamlines and all experimental chambers was just over $3 million, and the total operating cost from 1980 to 1999 was $2.6 million. From 1999 to 2009, under the direction of T.K. Sham and Walter Davidson, the total operating costs (from NSERC and NRC) were over $2 million (increasing to about $400,000 per year for several years as CLS began operating) as the success of CSRF facility became well known. As expected, the facility was initially dominated by users from Western; but a quick perusal of CSRF theses4 and publications* show that users from ten other Canadian universities and organizations contributed over 50 per cent of the theses since 1993. Not only was the quantity of CSRF publications remarkable for the amount of money spent, but the high quality of these publications is indicated by the large number of citations that many have generated. But most importantly, CSRF provided the first SR training and experience for a lot of present CLS employees (mostly in beamline design and construction) and users. For example, former CSRF employees at CLS in the last fifteen years include Kim Tan, Brian Yates, Yongfeng Hu, and Narayan Appathurai. Other CLS employees – Ian Coulthard, Jeff Cutler, Jigang Zhou, and De-Tong * Available from NRC Press or directly from [email protected]

The University of Western Ontario  39 Figure 8.  Some of the important Canadian synchrotron scientists in the quest for CLS. Clockwise from top left: Daryl Crozier, Adam Hitchcock, T.K. Sham, and Masoud Kasrai.

Jiang – did their PhD or post-doc work there as well, Emil Hallin from SAL was introduced to SR beamlines (see figure 9), and Tim May was a staff scientist at Aladdin. Many faculty members who are frequent users of CLS did their first SR research at CSRF. For example, Stephen Urquhart and John Tse, both at USask, did a lot of research at CSRF; and other Canadian faculty, such

40  The Canadian Light Source Figure 9.  CLS employees trained at the Aladdin synchrotron standing with the CSRF Grasshopper monochromator at CLS. Back row: Jigang Zhou, Tim May, Kim Tan, Ian Coulthard, Mike Bancroft, and Emil Hallin. Bottom row: Yongfeng Hu and Jeff Cutler. Missing: Brian Yates (see figures 6 and 7) and Narayan Appathurai.

as Adam Hitchcock, Ron Cavell, T.K. Sham, and Grant Henderson (University of Toronto), did most of their initial SR work there as well. The first publications from CLS resulted mostly from these researchers using the two soft x-ray beamlines that replaced the CSRF Grasshopper and SGM beamlines. As is usual for scientific and academic organizations, a few outstanding individuals (such as T.K. Sham, Masoud Kasrai, and Adam Hitchcock) contributed greatly to the success of CSRF with a large number of excellent publications. A few others, by contrast, published very little with comparable amounts of beamtime. In another example, a sophisticated microscope purchased with an NSERC grant of approximately $1 million for use initially at Aladdin resulted in no publications after four years. Other comparable instruments at Aladdin produced dozens of papers. Fortunately, such negative cases are very unusual: the vast majority of scientists work very intensely and publish effectively. It is extremely pleasing for me to see that the former CSRF students and users were critical to CLS’s development. I have to thank again the large number

The University of Western Ontario  41

of talented students, post-docs, senior scientists, and faculty who put great effort into the success of the remote CSRF facility. And all of us must thank the NRC and NSERC committees for their efforts, and especially the SRC staff and management at the Aladdin synchrotron for their outstanding co-operation. Certainly, this must be a model for international co-operation. Ed Rowe’s generous spirit was incredibly important for the Canadian group. Hopefully, our excellent research, which did not overlap with other research at the facility, enhanced the reputation of the SRC facility and generated more operating money for them. In the synchrotron area, the potential research is unlimited in scope, so direct competition is usually minimal. This leads to the co-operative support of most synchrotrons by other international synchrotrons.

4 Formation of the Canadian Institute for Synchrotron Radiation and competition between Western and USask, 1989–97

a)  Early Unsuccessful Efforts, 1989–95, and the Importance of NSERC By the late 1980s, there were at least fifty Canadian scientists using SR, ­primarily at second-generation US facilities. Many of them used the two ­Western beamlines in the soft x-ray region at CSRF in Madison; but others were using Brookhaven and Stanford in the hard x-ray region – for example, to obtain detailed structural information on proteins. And at that time, there were many new third-generation synchrotrons being planned in the United States (the ­Advanced Light Source [ALS] at Berkeley and the Advanced ­Photon Source [APS] at Argonne National Lab near Chicago), Germany, Sweden, ­England, and Japan. These third-generation sources were much more intense and brighter than the second-generation sources (such as Aladdin in Madison) due to the recent introduction of multi-polar magnets such as wigglers and undulators (see appendix 1). Many Canadian scientists began to think of yet another effort to fund a dedicated Canadian light source. In July 1989, Bruce Bigham at the AECL in Chalk River hosted a meeting of the accelerator physics and experimental user groups to identify the applications of SR and the Canadian users. Over fifty scientists in Canada that used SR abroad were identified. At that meeting, Bruce made the excellent suggestion that we create a corporation, the Canadian Institute for Synchrotron Radiation (CISR), to document all the Canadian synchrotron ­users and to rally further support, analogous to the Canadian Institute for ­Neutron Scattering, which had already been established to rally support for the neutron scattering researchers. Then, in January 1990, forty scientists from academia, government, and industry met in Ottawa at an NSERC-sponsored workshop. At the meeting, it was formally suggested that CISR be formed to coordinate Canadian activities. A constitution was ratified at the first annual general meeting in Toronto on September 15, 1990, and letters patent were received in D ­ ecember 1990. There were only three places in Canada with the

The Canadian Institute for Synchrotron Radiation  43

expertise to build the necessary electron accelerators: first, TRIUMF at UBC; second, AECL; and third, SAL. Representatives from the three labs (Bruce Bigham from AECL, Michael Craddock from TRIUMF, and Dennis Skopik and Les Dallin from SAL) were present at those first meetings. AECL had been undergoing many changes, and a Canadian synchrotron might give it an important new scientific direction. Indeed, Bruce Bigham was interested in such a synchrotron project; he also suggested that a new nuclear reactor was required in Canada to replace the then forty-year-old reactor that was badly needed for medical isotope production as well as other research such as neutron scattering for m ­ aterial analysis. I became the president of CISR, with the Board of Directors comprised of Daryl Crozier from Simon Fraser University, Ron Cavell from the University of Alberta, Adam Hitchcock from McMaster, T.K. Sham from Western, and Dennis Skopik from SAL. The Board of Trustees, with representatives from all the university and industrial members, was chaired by Adam Hitchcock. We rallied support from all over Canada, with individual memberships ($10 a year) and institutional memberships ($100 a year) paying travel expenses for outstanding international speakers at workshops and conferences. We had ten annual meetings from 1990 to 1999, and many workshops (of which more below). The membership grew quickly from about 60 individual members in 1991 to 220 individual members and 30 institutional members in 1997 when the final applications were submitted to the CFI for funding. Symposia at Canadian scientific society meetings were critical in growing ­interest and memberships. For example, Adam Hitchcock and I organized a fourday SR symposium, entitled “Chemical Applications of Synchrotron R ­ adiation,” at the annual Canadian Society for Chemistry (CSC) meetings in June 1991 in ­Hamilton. There were over fifty presentations, mostly from Canadian users. We then hosted symposia at most annual chemistry or physics meeting for the next seven years. The annual CISR meeting was also held at these conferences. Having identified most of the Canadian users and their scientific interests, in May 1992 an application for study funds for a Canadian synchrotron was submitted by CISR to Pardeep Ahluwalia, the head of NSERC’s C ­ ollaborative ­Research Initiatives program. The request was for about $900,000, with $750,000 for SAL and $150,000 for AECL. The SAL accelerator physics group (mostly Les Dallin at SAL) had already done a good deal of work on the ­designs; but these needed to be firmed up, and studies on superconducting magnets and c­ avities were needed from AECL. This proposal consisted of two large ­volumes  – a­pproximately 250 pages in total – and the two volumes were ­assembled at Western and SAL. I ­remember that NSERC put together an advisory committee; we met with them in Ottawa in the fall of 1992 in a review meeting, and again received no funding, as notified in early 1993. ­Pardeep Ahluwalia was our principal contact with NSERC for several years, and it became quite obvious to many of us that he felt that a Canadian synchrotron was not a high priority for Canadian science.

44  The Canadian Light Source

It became apparent to the CISR community by 1993 that only the SAL group was in a good position to build the new Canadian ring. The CLS project was absolutely essential to their survival, and they indeed had excellent technical and project management expertise (from UMA, the Saskatchewan engineering company) to build the electron accelerators and associated facility. This was a great strength for their application. However, they had very few SR users, and that was a large weakness. The only two USask scientists that had used synchrotron radiation were Wilson Quail in chemistry and Louis Delbaere in biochemistry. Both had used protein x-ray crystallography beamlines in the United States, Germany, and Japan. Moreover, there were still only these two users at USask in 1999 when CLS began construction. This in itself might have eliminated USask as the CLS site, because the local scientific base is usually key to the formation of any large scientific facility. In addition, some officials at the major science funder Industry Canada,* who were not keen on the project anyway, were not complimentary about the prospect of the facility going to Saskatoon. However, success would depend largely on fundraising, enormous intensity and energy from many people, coordinated efforts, political lobbying, timing, and luck. Luckily for USask and the whole project, the Liberal Party was elected in September 1993, with Jean Chrétien as prime minister, Paul Martin as m ­ inister of finance, and Ralph Goodale – from Regina – as minister of agriculture. Doug Richardson, a lawyer with Mckercher Mckercher and Whitmore in Saskatoon, the university’s law firm, had excellent contacts with Liberal Party members after a stint as parliamentary secretary in Ottawa in the 1980s. He did a lot of successful lobbying on behalf of USask and CLS. However, for the first three years of the new government, Paul Martin struggled greatly to get the large deficit under control, and there was no possibility of new federal funding until the deficit was tamed. As discussed below, Ralph Goodale managed to convince Western Economic Diversification to contribute to the SAL budget long enough to keep SAL afloat during this crucial time. But what suspense for SAL and its employees, and the rest of the Canadian synchrotron community! Meanwhile, by 1993, the presidents of NSERC – Peter Morand until 1995 (and his vice-president, Nigel Lloyd) and Tom Brzustowski after 1995 – were very supportive of such an initiative; and in 1994, NSERC formed a Committee on Materials Research Facilities (CMRF) in Canada chaired by David Bacon (Queen’s) to advise NSERC on what the priorities in this area were at the time. * Industry Canada (now entitled Innovation, Science and Economic Development Canada, or ISED), is a department of the Government of Canada. Its mandate is to foster a growing, competitive, and knowledge-based Canadian economy. With over 4,800 employees, it oversees all Canadian science-related organizations such as Western Economic Diversification, NSERC, and NRC.

The Canadian Institute for Synchrotron Radiation  45

Many outstanding Canadian and international scientists were engaged to help draft a long report and analysis. Indeed, the international co-operation brought to these efforts was truly amazing. It is important to emphasize that all these outstanding scientists donated their time (as they always do for any extra a­ ctivity) to these arduous efforts, obtaining just their travel expenses for many long days of work. The committee recommended that Canada (1) build a f­ acility to replace the National Research Universal (NRU) reactor at Chalk River for neutron beam research; (2) build a third-generation SR source; and (3) promote the establishment of “clusters” of equipment for materials research (such as SSW). This report arrived in late 1994, but there was no money anywhere to construct either a new nuclear reactor or the synchrotron. After the NSERC Council considered the three recommendations in October 1994, it decided to support all three; but because of the monetary situation (Canada was still running large deficits), NSERC could not act on the two large facilities. However, the above council mandated NSERC to co-operate with other stakeholders to further explore the building of the two large facilities; and in particular that “NSERC should initiate cooperative efforts ... to explore ... ­synchrotron radiation sources.”* While NSERC began to initiate efforts toward building a Canadian synchrotron facility, it still had no obvious way of funding the project. Indeed, the president of NSERC, Peter Morand, wrote me in late 1995 saying that he could see no way of funding the project for the next ten years. This letter did not decrease my enthusiasm or intensity in the least. However, it did not help me at ­Western with President Paul Davenport, who was not supportive of the project. But ­despite the negative financial picture, NSERC continued to help the ­synchrotron initiative, beginning with support for two workshops in November 1994 at USask and February 1995 at Western, with the goal of coming close to a final plan for the accelerators (USask) and the beamlines (Western). These workshops were ­organized on very short notice. For example, I wrote Bob McAlpine of NSERC on December 1, 1994, with a request of $6,000 to pay travel expenses for four international speakers invited to the February 1995 workshop. This small amount of funding (typical for academic workshops) was augmented by $3,000 from CISR and the CCP. The workshop included thirteen invited talks on beamlines and synchrotron science in some of the a­ reas that are important at CLS ­today: protein crystallography, medical imaging, XAFS for biology and materials science, lithography, and micromachining. It did not * As an aside, the nuclear reactor facility never did get off the ground for a number of reasons, and that proposal is now dead. This means that the nuclear isotope production for Canada and abroad from NRU will have to shift to other types of accelerators. CLS began an initiative in 2010 to produce such isotopes with electron accelerators, and CII, a private company, was established in 2015. Saskatoon then will host not just CLS, but an important part of the replacement for NRU.

46  The Canadian Light Source

include a lot of the ­well-established soft x-ray science at CSRF. These workshops resulted in a “­generic” plan for a Canadian SR facility, with the cost for construction of the complete synchrotron facility and ten beamlines being established at $130 million. The plan was to be made available to all potential host sites. It is important to emphasize here that senior NSERC personnel such as Bob McAlpine (and earlier Janet Halliwell, Isabelle Blaine, and Carmen Charette, to name a few) were critical in the 1980s and ’90s to the development of CLS. I think that most academic scientists take these dedicated civil servants for granted. But in my experience with over ten site visits for evaluation of major equipment and operating proposals, they have done a critical job very well: to ensure that both the scientists and the proposals are of the highest international standard and the budgets are reasonable. Perhaps their most important function is choosing an excellent committee chair (such as Boris Stoicheff for the initial CSRF proposal), followed by choosing a number of Canadian and international experts with little or no conflict of interest. Coordination of the travel for between six and eight committee members (often from Europe, the United States, and Canada) is a significant task as well. These NSERC personnel attend all meetings, take effective minutes for the committee, and help write ­voluminous reports and recommendations. The synchrotron experts ­donate their time for two or more days (including travel) to help their colleagues ­succeed, knowing that a successful new facility might compete with their own research. This model is used in most Western countries at least, and has to be the best example of free trade that I know of. Bob McAlpine must have spent a significant fraction of his time on the CLS file from 1994 to 2001, including coming to Saskatoon for CLS committee meetings when I was director. For example, the final 1994 CMRF report consisted of seventy-two pages of dense, single-spaced type – the result of an enormous effort by McAlpine and the committee. McAlpine also gave a lot of guidance on the overall process that NSERC envisaged would lead to a C ­ anadian synchrotron. The excellent co-operation from NSERC staff (and later CFI staff) was, and is, critical to the success of CLS. But there was still no obvious way forward for funding, and with the $2.5 million NSERC operating funding to SAL about to disappear by 1995, the prospects for SAL’s survival were looking bleak. Dennis Skopik and the ­Saskatchewan proponents approached Liberal cabinet member Ralph Goodale to obtain support from Western Economic Diversification (WED) for bridge funding to help SAL through to CLS. This funding came in two allotments, $500,000 for 1996–7 and $1.3 million for 1997–8 to cover the period before the CLS funding was allotted. Both WED and Goodale would be critical for CLS in the ensuing years, as we will see below. At about this same time, Dennis Skopik had sold the CLS concept to USask president George Ivany. Ivany had

The Canadian Institute for Synchrotron Radiation  47

been installed in that role in 1989, just as the university was entering a prolonged period of large provincial cuts. With a physics background, he appreciated the potential importance of CLS, and his focus on the project gave him an ­opportunity to build a legacy rather than continue in the total restraint mode. Dennis Skopik also sold the project successfully to the associate ­vice-president research, Dennis Johnson. A local lawyer, Doug Richardson (mentioned above), soon became involved as well. Richardson had a significant number of top-level Liberal contacts, including Paul Martin and Ralph Goodale. He began to lobby on behalf of USask, again in an incredibly high-risk environment with no possible money yet identified. Dennis Skopik also had strong support from the province, and it probably did not hurt that he played tennis at the same tennis club as the premier of Saskatchewan, Roy Romanow. Meanwhile, NSERC staff such as Bob McAlpine, working with CISR members, established a special committee to make recommendations on two points: (1) whether or not Canada should have a synchrotron, and (2) where it should be located. NSERC and CISR called for letters of intent from all ­Canadian ­universities in March 1995 outlining their proposals for the building of a ­Canadian synchrotron. These letters were to be presented by June 1995, and the final ­proposals by December 31, 1995. As was expected, the NSERC call for proposals resulted in two groups ­vying for CLS: the USask group (the so-called western Canadian proposal) and the Western group (the so-called eastern Canadian proposal). The competition ­between the two groups was intense but very friendly. For example, the two groups collaborated on the proposals, with a good deal of joint material, such as including the scientific user program in both. Competition is present in most human endeavours, and indeed is essential for obtaining the very best overall results. In the CLS’s case, competition pushed both groups to raise more funds, co-opt the most outstanding researchers in their region, design the best facility, and write the best proposal. But fortunately, the major personalities involved had great respect for each other, so that the loser was not devastated and completely uncooperative after the winner was announced. Peter ­MacKinnon12 points out the importance of such competition within any university. ­Remarkably, though, he had to convince some at USask of the ­importance of competition if USask was to become an outstanding university, not only locally but nationally and internationally.12 It is interesting to contrast now the efforts at the two universities vying to land the largest scientific project in Canada since TRIUMF in 1968. USask ­developed a coordinated effort at the highest administrative levels (USask president and associate vice-president research, the university board, the province, and the city) to secure the synchrotron project. Western, for its part, could not develop any coordinated effort at any of those levels, even with all of my previous experience in scientific fundraising.

48  The Canadian Light Source

b)  Competition between USask and Western, 1995–6 i)  The USask Effort As mentioned above, USask president George Ivany had a background in physics (a BSc in chemistry and physics and an MA in physics education). He was enthusiastic about this project, which he saw as a way of reviving the university’s reputation. He realized that losing SAL, the $2.5 million NSERC grant, and the talented people working toward the CLS goal would not be positive for USask. He and his senior colleagues were able to engage the USask board in a very difficult political climate. Fortunately, the chair of the board, Hal ­Wyatt, was also deeply involved, and Ivany and Wyatt managed to “stickhandle” the project through a university community that included a significant number of detractors, even in science and engineering. As usual in a university, a n ­ umber of faculty and departments were against hosting CLS because they thought they would lose funding if money was diverted to CLS. For example, the dean of engineering, Franco Berutti (later dean of engineering at Western), said at the time that USask was extremely strapped for resources just for the critical core of USask of operations, and should therefore forget about immediately jumping into a new opportunity. There would be substantial financial risk for any university in this situation, of course, but everyone knows that big risks have to be taken for big rewards. Fortunately, Dennis Skopik and his colleagues apparently convinced Premier Romanow and other politicians that this ­project could ­really put Saskatchewan on the map. Premier Romanow spoke eloquently about CLS as a “shining light to the world.” And yet others questioned the project with comments like, “Is this for the benefit of our children, or for Premier Romanow?” As of March 30, 1995, the USask Board of Governors had given approval in principle to the synchrotron proposal presented by Johnson and Skopik, so it was very easy to present a letter of intent for the facility immediately. Johnson and Skopik established the first coordinated committee to develop strategies for recruiting the project to USask. They met weekly in a boardroom near ­Johnson’s office in the rather austere surroundings of Kirk Hall, a postwar residence for returning vets. It was called the Synchrotron Steering Committee and, occasionally, the “Kitchen Cabinet” after another famous meeting. The committee was structured to have representation from the university, the provincial and city governments, the local private sector, and those with “influence” and connections to governments. Membership included Malcolm ­Sheppard (university controller), Dennis Skopik (director of SAL), Doug R ­ ichardson (Mckercher Mckercher and Whitmore), Doug Tastad (president, Innovation Place), Ron Woodward (president, Saskatchewan Research Council), Murray McLaughlin (president, Ag-West Biotetch Inc., later to ­ ­become deputy minister of economic development), Lorne Smith and Brian

The Canadian Institute for Synchrotron Radiation  49

Hansen (Saskatchewan Economic Development Department), and John ­Hyshka (of the Saskatoon Regional Economic Development Authority). Dick Batten, a lawyer from Mckercher Mckercher and Whitmore, attended some of these early meetings as well. There was obviously very strong representation from the Saskatchewan government, which hoped to oversee developments on the project. At this stage, there had been little or no discussion about ownership or management of the facility. It was entirely possible that it would be a federal science facility (for example, operated by NRC), but it could also be either a provincial or a USask facility. Brian Hansen became the most prominent Saskatchewan government employee on the CLS file, and he later advocated for provincial control and management of a large part of the project. Both USask and the Saskatchewan government were very interested in obtaining the facility. This would later lead to problems, but at least this meant the application appeared to be supported, and approved, at the highest levels. As chair of the so-called Kitchen Cabinet, Dennis Johnson reported on the group’s meetings to President Ivany and Vice-President-Administration Tony Whitworth; he also presented reports to the Deans’ Council of the University. In committee meetings, Doug Richardson spoke on government relations, Dennis Skopik on the national scientific community and CISR, and Malcolm Sheppard and Dick Batten on the possible management structure for CLS within USask. This committee discussed and plotted government and public relations strategies that became so important in bringing the city and province on side. It was recognized that provincial funding would need the support of the local population. Considerable effort was therefore made to ensure that the general public in Saskatoon and throughout Saskatchewan were aware of, and supported, the university’s pursuit of the synchrotron. In order to publicize the project, Dennis Johnson spoke to numerous service organizations in Saskatoon and to the Chambers of Commerce of Prince Albert and Moose Jaw, and even in Alberta. In Saskatoon, with the co-operation of the press, even the taxi drivers were aware of the project. These efforts certainly contributed to the financial support eventually provided by the city and province. The committee continued to meet through 1995 and 1996, with each person reporting progress in their area of influence and proposing preparations for the NSERC site visit in May 1996. It was during this period in early 1996 that the USask group obtained the lead CLS funding of $10.5 million from the Province and the $1.5 million from the city of Saskatoon. These amounts were increased greatly when more matching funding was required. These initial contributions clearly showed the commitment of the Saskatchewan community, and were worth far more than their monetary value. Indeed, they absolutely sealed the deal for the USask bid, partly because of the lack of financial commitment for Western’s proposal.

50  The Canadian Light Source

Meanwhile, Dennis Skopik and his colleagues developed the proposal in-house, with Les Dallin being critical for the machine design and Emil Hallin being critical for the building, labs, and beamlines. The politicking was perhaps just as important. Dennis Skopik, Doug Richardson, and Dennis Johnson travelled to Ottawa and Toronto to meet with government officials Ralph Goodale, Kevin Lynch (the deputy minister of Industry Canada), Donald Macdonald (the former minister of finance), NSERC, and various industry leaders. It was especially critical to get support from Industry Canada for the CLS project. The role Doug Richardson played from the outset to the conclusion of this project has probably been understated. Indeed, Paul Martin, the federal minister of finance, commented to me in a telephone interview in early 2010 that Doug’s role “was critical from my point of view.” He made all the early appointments with industry, politicians, and civil servants, and indeed, he still works on CLS-related projects today. Because Doug Richardson played such a pivotal role, it is important to look at his initial early contributions to our efforts to convince politicians that this project was of a very high priority and deserved strong financial support. The earliest correspondence available is a letter from Richardson to Ms. Pat ­Youzwa, then deputy minister of economic development, dated January 18, 1995. The letter makes the case for a Canadian synchrotron, describes SAL, and, finally, asks if Ms. Youzwa would meet with Doug, Dennis Skopik, and Dennis J­ ohnson. The text indicates he was fully involved even in 1994. A March 27, 1995, memo from Doug to Ivany, Skopik, and Johnson describes the March 24 meeting with Ms. Youwza. Doug was in Regina and was accompanied by David Miller of Hill and Knowlton. Doug knew David Miller from his days in Ottawa. The memo states that the meeting was called to work out “how Treasury Board and Cabinet will deal with the matter” (the request for a provincial contribution of $10 million toward CLS). It appears that Youzwa thought that the Saskatchewan government should agree in principle and not commit to funding until the federal government committed. That same day Doug met with Gary Aldridge, chief of staff to Premier ­Romanow. Doug reported that Aldridge said the government and premier were in total support of the project. Aldridge also told Doug that he should meet with the Honourable Ed Tchorzewski and the Honourable Janice ­MacKinnon, minister of finance, to insure that the key Treasury Board/cabinet people would give approval for our initiative. Throughout the early months of 1995, Doug wrote both provincial and federal officials – including Paul Martin – to ensure they were well informed about the project. While in Ottawa he met ­high-ranking ­officials such as Dr. Chaviva Hosek, senior policy advisor in the Prime ­Minister’s Office. In June, Doug made arrangements in Ottawa for D ­ ennis Skopik and Dennis Johnson to meet with several members of ­Parliament, ­including Jon Gerrard (Portage la Prairie), Reg Alcock (Winnipeg),

The Canadian Institute for Synchrotron Radiation  51

Dennis Mills (Toronto), and Ralph Goodale (Regina). During the same trip, Skopik and Johnson met with Arthur Carty, president of NRC, and Marc Lepage of the Medical Research Council (MRC; now the CIHR). As would be expected, these activities received full financial support from USask. Doug was also active in helping USask secure funding from the city of Saskatoon. Together with Brian Hansen, Skopik and Johnson met with ­ Marty Irwin, the city commissioner, who along with Mayor Henry Dayday and ­Alderman Peter McCann actively promoted the project. Eventually the ­Saskatoon City Council contributed $2.4 million to the project. ii)  The Western Effort Meanwhile at Western, there was little or no talk at the president’s or vice-president’s level of hosting CLS; and much greater and more fundamental difficulties than at USask arose when the synchrotron was mentioned. Some background to the Western effort should now be outlined. In 1993, a committee was established to choose a replacement for George Pedersen, who would be stepping down in July 1994 after ten years as president of Western. I talked to George about CLS in 1993–4, and I think he would have supported a ­Western initiative strongly. I was appointed to the Western presidential selection committee in September 1993. The ten-member committee consisted of five senior Western faculty and five appointees from the university’s Board of Governors. The committee was chaired by a very fine lawyer from the Western board, Claude Pensa. All committee members were conscientious and thoughtful; but I was surprised that I was the only faculty representative from science, engineering, or medicine. Moreover, only one of the five appointees from the board, a London lawyer, had a strong background in those research-intensive areas. As with most senior university appointments in Canada, a consultant (or headhunter) had been hired to recruit excellent candidates and raise letters of reference for the short-listed candidates. In early meetings, the desired role and direction of the university was discussed at some length. The discussion was partly focused on whether Western should be mainly a southern Ontario liberal arts institution with some strong research areas, or whether it should aspire to be a research-intensive university of international stature. Partly because of the membership of the committee, there was no consensus about which of these two directions Western should take. A few applications came from the broad science area, along with others from the social sciences, such as Paul Davenport, an economist whose one-term contract as president of the University of Alberta had not been renewed. After a few candidates were i­nterviewed, I mentioned at one of the last meetings, held in early 1994, that there was a significant possibility of CLS, the science project of a generation in Canada, coming to Western. As expected for a committee with little

52  The Canadian Light Source

science representation, there was scant appreciation of, and no support for, the ­synchrotron project at Western. The consultant produced several dozen very positive letters of reference for Dr. Davenport, but many fewer and less positive letters were produced for the other short-listed candidates. Based on these letters of reference – and of course the interviews – Davenport was selected as president. As discussed below, this one decision was a major negative factor for the Western-based CLS application. Again, Peter MacKinnon in his book12 points out how important the committee selection process, committee membership, and consultants are to such a presidential selection. Quoting from an American university task force, he mentions that there can be “excessive reliance on a presidential search consultant to carry out the Board’s responsibilities.” Paul Davenport began his tenure as president in July 1994. Over his fifteen years in that role, he was very dedicated and very effective at boosting the undergraduate profile of Western. An excellent speaker, he could give a convincing ten-minute talk on virtually any area of research at Western. But I realized in 1994 that there was a very small possibility of convincing a non-scientist that CLS would be a reasonable financial risk and good for much of the university, especially the sciences, engineering, and medicine. The vice-president research, Glen Caldwell (who had been at USask for close to thirty years before coming to Western), was in support of the CLS initiative and helped me greatly, as did his successor Bill Bridger (who assumed the role in January 1996). Because Davenport arrived at Western in July 1994, Caldwell and I did not approach him until early 1995 for help with the Western application to NSERC. Unlike USask, which would lose the approximately thirty-five SAL employees if it did not win the CLS project, Western would not lose any large group of researchers if it did not land CLS. It was possible that I could have ended up at USask, along with T.K. Sham and Masoud Kasrai. I had planned to resign as chair of the chemistry department in May 1995 after eight years in order to concentrate nearly full-time on the CLS application due December 1, 1995. I still had a research group of over ten students and postdocs, and I was still travelling to Madison regularly to oversee CSRF operations. Moreover, from January to April 1995 I was still teaching an undergraduate course, and I was very busy with a faculty member’s problems in the chemistry department. Both the USask and eastern Canadian groups worked very hard for six months to produce enormously impressive proposals – both six volumes and over seven hundred pages long. I was able to complete this proposal with the aid of only $50,000 from the deans of science and medicine. I hired a machine physicist, Mike Green, from the Aladdin synchrotron in ­Madison, on contract. I also hired a full-time post-doc, Arthur Bzowski (from T.K. Sham’s group), to help write the proposal; and I also hired my former chair, Howard Clark, who had recently retired as president of Dalhousie University, to canvas

The Canadian Institute for Synchrotron Radiation  53

industrial interest in eastern Canada. Adam Hitchcock from ­McMaster was ­always helpful with the proposal writing. Jeff Cutler, a former graduate student, came back to Western in the summer of 1995 as a post-doc, and he also worked effectively in developing the Western proposal. Three committees were organized, and the above-mentioned consultants hired, by August 1995. The two deans wrote a very strong letter to President Davenport on July 21, 1995, giving very strong support to the synchrotron initiative at Western and committing the $50,000. The first meeting of the Western synchrotron group was held on August 1, 1995. The Development/­Management Committee was chaired by me along with T.K. Sham, Martin Stillman, Adam Hitchcock from McMaster University, and Yong Kang, the dean of science. The Outreach Committee, chaired by Howard Clark, included several senior administrators and previous administrators from Western, among them Glen Caldwell (vice-president research), Al Adlington (former vice-president ­finance), Don Hayden (associate dean of science), and Paul Harding (associate dean of medicine). The Industrial Strategy Committee was chaired by ­Stewart McIntyre, the director of SSW, and included several SSW employees such as Tim Pope and Mary-JaneWalzak, and the main post-doc working on the project, Arthur Bzowski. There were only three significant consulting salaries paid for by the $50,000, to Howard Clark, post-doc Arthur Bzowski, and Mike Green. The political part of our lobbying efforts (for example, politicking at Queen’s Park) was ceded to the senior administrators at Western, such as Glen Caldwell, Paul Davenport, and the London manager of business development, John ­Winston. I never did get any news back from the “fundraisers,” and I do not even know if they had official visits to Queen’s Park. But there were no “highlevel” meetings with President Davenport, the chair of the Board Governors, or city/provincial administrators. I made several visits to Ontario universities and academic groups to gain support in Ontario. Although there was good support from some universities, such as McMaster and Guelph, a meeting with the vice-president research at the University of Toronto resulted in no central support from that institution. Fortunately, Toronto’s dean of medicine, Chris Yip, did support the proposal because there were several very active protein crystallographers in his faculty. SR facilities have been extremely important for the many protein crystallographers in Ontario and Canada since the early 1990s; but another meeting in December 1995 with a Canadian protein crystallography group did not result in enthusiastic support for a Canadian synchrotron facility. Many members of this group were already successfully using facilities in the United States and Japan, and they did not think that a Canadian facility could compete with the ones they were using. This “put-down” of Canadian abilities and achievement is of course common in many areas, and is nearly always incorrect. The protein crystallography beamlines at CLS have now been widely and very successfully

54  The Canadian Light Source

used by many Canadian and US groups (see chapter 8 in this volume). And yet, until recently some of the older and more established groups nonetheless kept using US facilities instead of CLS. There was only one more meeting of the above committees, on November 23, 1995, to discuss the proposal due on December 1 of that year. Even at this late date, the Western board had not been notified of the project, and there was not a dime on the table from either the province, the city, or Western. A significant part of the problem lay with the newly elected government of Mike Harris, which was cutting nearly all budgets in Ontario for the next six months – e­ xactly the wrong time for our proposal. In contrast, USask, whose Steering Committee (or Kitchen Cabinet) was meeting every week near Johnson’s office, quickly raised $38 million from the university, province, city, etc. (of which more below). Compared to the budget-slashing then going on at Queen’s Park, the Saskatchewan provincial budget was in reasonable shape after a lot of financial problems in the 1980s and early ’90s. Diane Cunningham, a well-known Ontario MPP, lived down the street from me, and I knew her well; but it was not even possible to get her involved, as was the case with the city of London’s manager of business development, John Winston. Most provincial spending was being cut, and the timing was incredibly bad. But even with provincial and municipal support, it is highly unlikely that the Western proposal would have been successful, because President Davenport was just not supportive, and such support was essential. Glen Caldwell and I met about once a month with President Davenport, beginning in May 1995 until the submission in December 1995. It was obvious that he was not g­ oing to do anything significant to help raise funds for the CLS project. For example, he did not take the proposal to the Western board for any kind of support. By S­ eptember 1995, I wondered why I was doing this huge and complicated proposal, knowing that there was little or no chance for success at Western without financial commitments from the university. I rationalized this effort by telling myself that I was doing it to advance the overall cause of CLS, mainly by pushing USask to greater efforts through the competition. I was also helping the soon-to-be laidoff employees at USask, along with a number of my former students who would be leaders at the new facility. Meanwhile, I had small group meetings every two weeks with the major players developing the Western proposal (mostly Arthur Bzowski, Jeff Cutler, and Adam Hitchcock). And yet, in none of these meetings were any of the senior Western administration or board members or municipal or provincial officials present. There was also no one from Western to even help write the proposal. Contrast that to the situation today, where there is a lot of help from scientists in the administration at Western (and most other universities) hired to help write grant proposals that, compared to the synchrotron proposal, are much smaller in scope and financial requirements.

The Canadian Institute for Synchrotron Radiation  55

So, although there were difficulties at USask, they were able to generate the necessary high-level support, which resulted, in turn, in the necessary large financial support. In contrast, at Western it was not possible to generate any high-level support – a situation that resulted, of course, in no financial support. Nonetheless, I took the six-volume, 720-page proposal to Davenport at the end of November 1995 with enormous pride and satisfaction. This proposal was the largest and best of all my large volume of research and grant writing – including my 250-page book18 and many equipment and operating grants totalling over $10 million (for example, all the initial grants for SSW equipment as well as for all the grants for CSRF in Madison over twenty years). My post-docs, Jeff Cutler and Arthur Bzowski, energetically and effectively supported by Adam Hitchcock from McMaster, had put in long hours to get this proposal finished. Davenport’s words to me and his vice-president administration, Peter Mercer, still ring in my ears over twenty years later: “Mark my words, this project will be a bigger financial sink-hole than the research park.” As an aside, the research park at Western had not been managed well for years, and had at least $10 million of debt that the university had to pick up just at this time. But to put these numbers in perspective, Western spent over $180 million on buildings during Davenport’s fifteen-year tenure at the university. Even with the very considerable provincial funding cuts, I speculate that another president at Western would have promised at least $20 million to the project – an amount that might have secured CLS at Western. I was of course very discouraged by Davenport’s comments, but I had to continue being positive for the hopeful benefit of the project. Davenport gave the whole proposal to Peter Mercer to read and evaluate, ignoring his vice-president research, Glen Caldwell, who was supportive of the project. There was little chance that a non-scientist could comprehend the very complicated scientific and financial proposal. iii)  NSERC Site Visits in May 1996, and the NSERC Decision The proposals from Western and USask were first vetted by CISR and forwarded to NSERC on February 1, 1996, in order to be reviewed by an ­international peer review committee. No more interest was shown by Paul Davenport at W ­ estern in the months leading to these site reviews, which were conducted by the A ­ dvisory Committee on Site Selection for the C ­ anadian Light Source Proposal. In great contrast, George Ivany at USask was still involved through his regular briefings from the Kitchen Cabinet meetings. The site selection committee was chaired by Alex McAuley from the ­University of Victoria, and its members were David Pink from Saint Francis Xavier ­University and four outstanding international scientists and highlevel administrators (three in the synchrotron area): Ingolf Lindau from

56  The Canadian Light Source

Lund, Sweden, John Madey from Duke University, Stephen Milton from the APS outside Chicago, and Gwyn Williams from the National ­Synchrotron Light Source at Brookhaven, New York. Most of these scientists were not well known to any of the Western or USask scientists. It is worth noting, again, the importance of international scientists to the success of any C ­ anadian project. These talented people donate a very large amount of their valuable time to help their international colleagues, many of whom they have never met. It is also important to emphasize again the important role of Bob McAlpine from NSERC at these meetings. The committee based its decision on the multi-volume proposals, written reports from more external reviewers, and visits to each proposed site: Western on May 6 and 7 and USask on May 8 and 9, 1996. The NSERC Site Review Committee spent one very busy day (May 6, from 8 a.m. to 9 p.m.) plus the morning of May 7 at Windermere Manor, at the ­Western research park, just off the Western campus. The eastern ­Canadian proposal out of Western was supported in a number of ways by eleven other eastern Canadian universities, including Waterloo, Windsor, McGill, Laval, Sherbrooke, McMaster, Guelph, Toronto, Ottawa, Quebec, and Montreal. They all sent strong letters of support; and McMaster, Guelph, Waterloo, and Toronto sent senior representatives – Ron Childs, Bob McCrindle, ­Carolynn Hansson, and Chris Yip, respectively – to talk about their present academic support for the eastern Canadian proposal, along with possible new academic positions in the synchrotron area at their universities. In ­addition, the proposal was supported by letters from eighteen major eastern ­Canadian industries; and 3M, the Protein Engineering Centre of E ­ xcellence at the U ­ niversity of Toronto, and Kurt Lesker Canada sent senior representatives Al Mills, George Connell, and Paul Robinson, respectively, to talk about their ­support of synchrotron research in eastern Canada. President Paul ­Davenport, Vice-President Research Bill Bridger, and Dean of Science C.Y. Kang ­represented Western, and Davenport gave the welcoming talk, stating the support from Western – which included six faculty positions and land in the research park for the building. However, there was no direct fi ­ nancial support from Western, and no plan to hire a number of the key scientists and technologists from SAL. There were also representatives from the three levels of government: John Winston, economic development officer for the city of London; Dianne ­Cunningham, MPP for London West; and Sue Barnes, MP for London. While John Winston sat through the entire proceedings, the latter two attended only the dinner. None had any money to put on the table. To his credit, John realized the importance of raising money immediately, and expedited a $3 million contribution from the city a week after the meeting – too little and too late. The lack of any financial support had to result in dismal failure.

The Canadian Institute for Synchrotron Radiation  57

In addition, Howard Clark and George Connell (former president of ­Western and later president of the University of Toronto who had funded the CSRF effort in 1982) were also present, with Howard Clark giving a talk in which he provided an overview of industrial support for CLS. Indeed, George Connell flew in from Halifax to support this proposal on his own dime. I then gave the case for building CLS in London. This case centred on four major facts. First, over 65 per cent of the existing SR users were in eastern ­Canada, with 48 per cent in Ontario and 17 per cent in Quebec (this compared to less than 5 per cent in Saskatchewan and 20 per cent in all four Western provinces). Second, the vast majority of the industries that would use CLS were again in eastern Canada. Third, the majority of the beamline and experimental expertise was in eastern Canada. And fourth, the management experience from large groups (such as SSW and ISW) was available to run the synchrotron for both academic and industrial users. Mike Green, the main synchrotron designer from Aladdin, gave a talk on the design of the synchrotron, which was rather similar to that designed by Les Dallin at USask. He also presented a summary of the proposed costs of the physical facility. The total cost of $127.4 million for the building, ring, and beamlines was substantially higher than the western Canadian proposal at the time ($77.4 million). The operating cost of approximately $25 million for five years was similar in both proposals. But again, the eastern Canadian proposal had no money on the table, whereas the western Canadian proposal had over $14 million in direct contributions and $22.7 million in in-kind contributions from the SAL facility. There followed a number of talks from several of the major Canadian synchrotron users on the need for the Canadian synchrotron; these were delivered by T.K. Sham and Mike Fleet from Western, Adam Hitchcock from McMaster, Emil Pai from the University of Toronto, and Tom Ellis from the University of Montreal. Stewart McIntyre, director of the highly successful SSW, gave a talk on the unique combination of academic and industrial research at SSW, which had generated more than $1 million in industrial revenue for over ten years in a row. The approach at SSW could be partially used at the new synchrotron. Don Hayden, director of the Western research park, then described the proposed serviced site for the synchrotron. Visits to the surface and interface labs at Western (SSW and ISW) showed the university’s unique strength in a prime synchrotron area, as well as the technical expertise in beamline development at ISW. These beamlines transported charged protons and ions rather than electrons or photons, but technically were very similar to synchrotron beamlines in many ways. Many posters at the ISW laboratory demonstrated the great scientific ­expertise in SR research at Western and in eastern Canada more broadly. The dinner that night included all the speakers as well as the two provincial and federal politicians, Dianne Cunningham and Sue Barnes, respectively. The whole meeting at Western was a great embarrassment for me. There was

58  The Canadian Light Source

no significant plan to hire the USask professional and technical staff (this was ­essential for the Western effort) from either the president or the dean of science, and no money promised from the province, the city, or Western. John Winston helped to raise $3 million in a week from the city of London. The province also became active and had several people review our proposal, but no money was forthcoming, as we learned in a letter of May 30, 1996, from William Saunderson, the Ontario minister of economic development. Such support would have been too late anyway. Meanwhile, on May 7, the NSERC committee landed in Saskatoon in a snow storm, where they were soon greeted by the provincial premier, the chair of the USask board, and many provincial and WED officials. The NSERC/CISR-sponsored site visit to USask occurred May 8–10, 1996, with Professor A. McAuley from the University of Victoria as chair of the peer ­review committee. The opening session was held at Innovation Place, the ­university-associated research park adjacent to the SAL facility. Dennis Johnson welcomed the visitors because President George Ivany was attending the funeral of his father. To put the university’s proposal into context, Dennis provided an overview of the university and the vibrant research and development community that surrounded it, including Innovation Place, ­Canada’s most successful university-associated research park. This community included the NRC’s Plant Biotechnology Institute, the federal Agriculture and Agri-Food Canada Research Centre, the federal Health of Animals Laboratory, and the National Hydrology Research ­Centre. ­Saskatoon had become the leading Canadian centre of agriculture-related ­biotechnology, dubbed by some “chlorophyll valley north.” The USask Vaccine and Infectious D ­ isease Organization (VIDO) was a highly successful, wholly university-owned research facility with numerous patents and royalty agreements. VIDO had its own Board of Directors, eventually serving as a model for governance of CLS. Innovation Place was also home to a number of software and communications companies, including SED Systems, a spin-off company from USask that provided research and development in the satellite communications sector. Dennis Johnson then outlined the long history of subatomic and accelerator physics at USask, noting that seven existing faculty positions were in this area. He also indicated that the president had promised six new faculty synchrotron user positions if the proposal was successful. He pointed out that SAL staff had the talent and expertise to build a synchrotron ring – as demonstrated by their design and construction of EROS – as well as the ability to manage a successful research facility over many years. He then spoke of how the Saskatchewan effort was well down the mobilization pathway, and the USask group had been striving for total community ­involvement in recruiting the project. Indeed, there was significant support

The Canadian Institute for Synchrotron Radiation  59

from the Saskatoon business community. The efforts of the Steering Committee (the Kitchen Cabinet) had resulted in commitments of $10.5 million from the province of Saskatchewan, $1.2 million from the city of Saskatoon, and $2 million from the SaskPower Corporation. Additionally, the University of Alberta had committed $300,000 for beamline development. Also, USask was contributing equipment, such as the linear accelerator and land, which together were valued at $22.7 million. The $10.5 million commitment from the province occurred after Dennis Skopik, Dennis Johnson, and Doug Richardson met with Pat Youzwa and then Frank Hart, both deputy ministers of economic development in the Romanow government. The $1.2 million capital commitment from the city of Saskatoon came after a study of water utilization at the SAL facility and the projected water use by a synchrotron facility. After consideration of the study, City Council voted to award the university the funding based on the costs to construct a water-cooling facility as part of the synchrotron project. This decision resulted very likely from the widespread publicity the project had generated in the ­Saskatoon community, the active involvement of an alderman, Peter McCann, and a very forward-thinking City Council chaired by Mayor Henry Dayday. The $2 million from SaskPower was largely negotiated by Doug ­Richardson in meetings with Jack Messer, the company’s president, and Messer’s vicepresidents. This contribution was made in recognition of the university’s status as a large consumer and of the additional electrical power usage that the synchrotron would generate in future years. John Hyshka of the Saskatoon Regional Economic Development A ­ uthority spoke of “the Saskatoon advantage,” comparing the wage rates, water, and n ­ atural gas power-consumption costs in Saskatoon compared to those in ­Ontario. Doug Tastad, president of Innovation Place, outlined the possible role of the research park in the synchrotron project and the park’s role in economic ­development more broadly. A representative of Nesbitt Burns examined funding alternatives, which would spread the federal government’s funding obligation over a term approaching the life of the facility. This involved establishing a trust that would borrow funds in Canadian capital markets equivalent to the federal government’s contribution and repay the funds, with interest, over time from an annual federal payment. The UMA engineering firm had a history of working with SAL staff, beginning in the early 1960s with the design of the initial building and lasting through three expansions and the construction of the EROS facility (see chapter 2, s­ ection c). UMA had also gained experience working with TRIUMF in ­Vancouver over a period of ten years. Barry Hawkins of UMA gave a presentation on project management, cost control, operating costs, and building design details.

60  The Canadian Light Source

After a tour of SAL, the site visitors were entertained with cocktails and d ­ inner at the Bessborough Hotel. On the morning of May 9, Dennis Skopik presented an overview of the USask proposal; this was followed by presentations on the scientific justification and applications of SR by Louis Delbaere, Ron Cavell, Marie Fraser, Lee Groat, Tom Tiedje, and Daryl Crozier – all scientists from western Canadian universities. The Universities of Alberta, Calgary, ­British Columbia, Victoria, and Simon Fraser University had been visited by Dennis Skopik and Dennis Johnson to develop support for the USask proposal in the fall and winter of 1995–6, paid for in part by the Saskatchewan government. The afternoon session involved discussion of the synchrotron ring philosophy (Roger Servranckx), its technical details including lattices (Les Dallin), radiofrequency systems (Mark Silzer), controls (B.E. Norum), and beamlines (Ron Cavell). Jack Bergstrom spoke on the lessons SAL staff had learned from the design and construction of the SAL prototype, EROS. The afternoon ended with a talk from Dennis Skopik on a transition plan for SAL and Dennis ­Johnson’s summary statement on the USask proposal. On the morning of May 10, the site visitors attended a working breakfast at the Saskatoon Club hosted by Premier Romanow. The selection committee held a meeting that afternoon at the Bessborough Hotel to consider the two proposals. On June 1, the McAuley report was released. It recommended that the ­ facility be funded at USask. On June 1 and 2, the CISR Board of Trustees met in ­Saskatoon to consider and endorse the report and to congratulate the ­USask team for winning the competition. Adam Hitchcock and I (chair of the CISR Board of Trustees and CISR president, respectively) wrote Dr. Tom ­Brzustowski, president of NSERC, to inform him that the NSERC-sponsored selection committee had made two recommendations: (1) that a Canadian national synchrotron ­facility be constructed as soon as possible, and (2) that such a facility be ­located at USask. As support for the USask application, the site selection ­report outlined the large financial support, the SAL equipment and expertise, the ­excellent project management support from UMA, and the proximity of a very successful research park. I was, of course, deeply disappointed that the project was not coming to Western; but I knew that with no financial support at Western, the committee could make no other decision. I certainly was pleased that the project would go ahead somewhere in Canada. For me, Saskatoon was the second best place in Canada – both because I had been born in Saskatoon and because I admired many colleagues in Saskatoon, such as Dennis Skopik, Dennis Johnson, Les Dallin, and Doug Richardson. On June 3, Premier Romanow, Dennis Johnson, and City Councillor Peter McCann spoke at a press conference announcing the award of the CLS project to USask.

The Canadian Institute for Synchrotron Radiation  61

In an appended executive summary, the total cost of the facility was budgeted at $115,856,000, with funding in place listed at $38,435,000 (summarized below); this resulted in a need for an additional $77,421,000. Also, at least $8 million was required to begin operation of the facility. At its meeting of June 12, 1996, the NSERC Council endorsed the two ­recommendations. But remember that there was still no mechanism for funding this proposal. Moreover, on the many trips made to Ottawa to lobby for funding, Industry Canada made it clear that it was not in favour. Indeed, an Industry Canada official told Doug Richardson and a USask group, “Not in my lifetime will this project be built in Saskatoon”! Formal organization charts were constructed for the overall project and its management, but the roles of CISR, the province, and the university were not defined. And the question ­remained: Where was all that money coming from and who was going to own and control the facility? Back in London, the London Free Press showed clever cartoons (see figure 10) on May 23 and June 5, 1996, depicting the CLS project going to Saskatoon. The editorials on those days were not enlightened. For example, the May 23 editorial stated: “As a city and province, we simply cannot continue to pump money into projects on such vague hopes of future rewards.” A separate, a­ lbeit well-intentioned, comment by a member of Western’s senior administration was especially annoying, and reflected the total lack of perspective on the ­nature and difficulties of large collaborative projects: “I’m very sorry that you were not successful, and I look forward to your next great application”! Of course, there would never be a similar application in my career, and the odds were small that anyone would write and assemble that kind of an ­application for many decades. My former very supportive dean of science, John Bancroft, wrote a critical letter in November 1996 to the Western newspaper outlining the enormous amount of time and effort expended over the previous fifteen years to build up expertise at Western to support the university’s CLS application. In it, he wrote: “Currently, over $20M of equipment is now in SSW, ISW, and CSRF and these units, with many talented and innovative faculty, staff and students, and research incomes of over $4.5M per year, have about 200 industrial and academic users per year and produce over 100 scientific publications per year. They constitute the best facilities of their kind in North ­America.” Of course the major problem at Western was the lack of support from the university administration for the Canadian synchrotron. I immediately got on board with the USask proposal, and it was decided that a joint USask-Western proposal would now be presented to NSERC. A total of $38.5 million had been raised, with $1.2 million from the city of Saskatoon, $10.5 million from the province of Saskatchewan, $4.0 million from WED, and $22.8 million from the SAL facility as an in-kind contribution.

62  The Canadian Light Source Figure 10.  London Free Press cartoons, May 23, 1996 (top), and June 5, 1996 (bottom).

The Canadian Institute for Synchrotron Radiation  63

c) June 1996–March 1997 and Control Issues with the Saskatchewan Provincial Government i) The Joint USask-Western Proposal to NSERC and the New USask Executive Committee: The Struggle for Funding Control for the project was now concentrated entirely in Saskatoon, but there were still two fundamental questions to be answered: Who would fund this facility? And who was going to own and manage it? The financing was the greatest issue because SAL was quickly running out of its interim money from WED. Also, there was a looming competition between the university group and a provincial government bureaucrat, Brian Hansen. He had no expertise in the synchrotron area; but, acting on behalf of the province, he apparently wanted the province to own and manage at least part of the facility. The first step in the process was to submit a joint proposal to NSERC by August 1996; this would combine all information from the USask and ­Western proposals into a single document, an idea that the eastern community out of Western, CISR, and I affirmed immediately. It is not usual for a losing bid to immediately endorse the winner, but this had to be the pragmatic decision for the benefit of the entire Canadian synchrotron community. Several meetings at Western and USask and many phone calls between the western and eastern teams firmed up this proposal, and it was submitted by Dennis Skopik and his team by the end of August. The three CSRF beamlines at Aladdin in ­Madison, Wisconsin formally owned by Western would be moved to CLS in Saskatoon, and probably modified or replaced. Bridge funding to retain SAL staff was being negotiated with the federal and provincial governments. In ­order to review the status of the project, a group of NSERC Council members headed by NSERC president Tom Brzustowski was scheduled to visit the USask campus on ­October 8, 1996. After endorsing the project, NSERC awarded SAL $500,000 in interim funding, which allowed SAL to complete previous salary and other obligations while developing the CLS project. The USask Kitchen Cabinet continued to meet after the May 8, 1996, NSERC peer review to evaluate the site visit to USask. Dennis Johnson’s last minutes from this committee are on August 14, 1996, and the combined USask/Western proposal was part of the agenda. The idea of a CLS office at Innovation Place, perhaps controlled by Government of Saskatchewan representatives, started to creep into the items to be covered. The last meeting of the Kitchen Cabinet was held on September 10, 1996. Just before that, Hansen and Ivany organized a new, higher-profile Executive Committee, which met for the first time on September 3, 1996, in the president’s office; the meeting was chaired by President Ivany and consisted of Tony Whitworth (vice-president finance, USask), Brian Hansen, Doug ­Richardson, provincial representatives David Dombowsky and Jim Yuel (both

64  The Canadian Light Source

S­ askatchewan businessmen), Jane Horachek (executive assistant to Georgette Sheridan, MP for Saskatoon–Humboldt ), Dennis Skopik, and Dennis Johnson. Johnson presumes this committee was established through conversations between President Ivany, Brian Hansen, and Georgette Sheridan. It was obvious that there was almost equal representation from the original USask group, whose members did all the work to get the facility allotted to Saskatoon, and the provincial government representatives, who undoubtedly wanted to help get the project funded and establish some government control in return for the province’s large financial commitment to the project. It is reasonable that the biggest funder of the facility at that time (the Saskatchewan government) might want to have important input into facility plans. Also quite reasonably, they would be very interested in any potential industrial revenue that the facility could produce for the benefit of Saskatchewan. The potential problem was that none of the provincial representatives had any scientific background; and they had very little knowledge of synchrotrons or how much industrial revenue could be generated from the project. Whether they liked it or not, they had to rely on the original group that worked so hard for years to land the facility in Saskatoon. Indeed, they wanted to create a niche for government involvement with very little consultation. Moreover, this was destined to be a national facility, and too much provincial control of such a facility would probably not have been reasonable. It is also possible that the provincial officials had little confidence in the university’s ability to secure funding for the project. ii)  Creation of a Team Canada Committee and the New CLS Office On July 11, 1996, after the nascent CLS project had been awarded to USask, President Ivany wrote the Honourable D. Lingenfelter, the provincial ­minister for economic development in Saskatchewan, seeking interim funds to support the local task force developing the synchrotron proposal. In this letter, ­President Ivany noted that the federal government had not made a final decision to build and operate such a facility. The interim funding would be used to establish a CLS office at Innovation Place, across the road from SAL in the USask research park, and to support travel for university-based scientists who would be participating in focus group meetings to recruit support for the project. Funds were also sought for seconding Dennis Skopik to the project on a full-time basis by supplying funds for a sessional lecturer to replace him in the Department of Physics. This help for Dennis was long overdue. The plan was to operate the CLS office for one year. The letter ended by indicating that Dennis Skopik and Dennis Johnson would be pleased to work with provincial officials to establish the CLS office. Lingenfelter responded on July 31, 1996, indicating the provincial cabinet had confirmed support for a series of strategic initiatives led by the Department of Economic Development.

The Canadian Institute for Synchrotron Radiation  65

On August 15, 1996, Brian Hansen, the provincial representative from the Department of Economic Development, sent President Ivany a thirteen-page fax containing a CLS Strategy and Action Plan (referred to below as the “­action plan”) indicating three major elements to an overall strategy: (1) the creation of a CLS Team Canada group; (2) establishment of the Office of CLS; and (3) the CLS Capital Campaign. The action plan outlined a course of action for CLS a­ lmost independent of USask. There had been no apparent discussion of the plan with the USask team. Many of the proposed activities were already ­ongoing under Dennis Skopik, Doug Richardson, Dennis Johnson, and the CISR academic community. Indeed, the CLS office was formally established by Brian Hansen in the ­research park in mid-August 1996. Hansen invited Skopik and Johnson to see the new office. There was a central office for Hansen, and one on each side for Skopik and Johnson. Was someone with no science background or expertise going to lead the project? This was an inappropriate takeover attempt: it was not going to build the type of teamwork required to fundraise for the project. The interactions with the government group caused Skopik and Johnson a lot of stress, partly because the government group did not understand the facility or the recommended enhancements made by the McAuley committee. Skopik, avoiding direct contact, wrote the government group to suggest they reread the McAuley report. Not surprisingly, Skopik and Johnson never occupied those spaces in the CLS office. By this time, Brian Hansen had become an assistant deputy minister of economic development. Thus began a struggle for control of the project, and this struggle was a vexing problem for all involved for at least the next two years. As an aside, a very similar situation happened in Australia, with the Victoria government bureaucrats in Melbourne actually taking total control of the Australian synchrotron with negative effects for many years. The history of this facility is still to be written. CLS Team Canada was to consist of industry CEOs, academic leaders, and “specific scientists.” Its mission was “to develop the argument for a Canadian facility in the context of enhancing Canada’s competiveness abroad through the use of advanced science and technology.” The action plan stated that the team of industrial and academic leaders would elevate the debate into the ­national arena as well. But, as it evolved, the critical statements in the above plan could be summarized by the following statement from Hansen: “Prior to the development of this strategy, the Province had little direct influence on the major ­elements of this project. In point of fact, the negotiations on behalf of this p ­ roject were left to a committee of the University and Saskatoon business organizations. Recent representations to the Premier suggested the Province contribute an additional $25M to the $10.5M already committed. The provincial position would only be

66  The Canadian Light Source

weakened by future, more public representations for additional contributions to the project by Saskatchewan.”* The committee referred to above was, of course, the same Kitchen Cabinet that had carried the project forward successfully through the first and most important step – the award of the project to USask by the NSERC-sponsored peer review committee. As indicated previously, the Kitchen Cabinet had already raised $38 million toward the project. Also, there was significant qualified government representation on the original committee from the beginning. Furthermore, strategies were being developed (and were actively underway) by that committee to recruit support from the academic community and the private and public sectors, as well as to build support in Ottawa. Many academics would have considered this type of provincial involvement as direct interference in the traditional autonomy of the university. In Saskatchewan, there had only been a few instances when provincial governments had been accused of disrupting this autonomy, as outlined in Michael Hayden’s book.10 (For more background on this issue, refer as well to Peter MacKinnon’s book.12) The Hansen action plan also stated that, “Through engaging a CLS Team Canada group, the Saskatchewan position can be formally endorsed and the CLS project considered in the context of the overall socio-economic benefits to Canada.” Team Canada was to consist of people from a large range of ­Canadian corporations, such as Nortel, Monsanto, Abbott Pharmaceuticals, Inco, and Cameco. Formation of the team was to be facilitated by Brian Hansen, ­David Dombowsky, Jim Yuel, Dr. George Ivany, and Doug Richardson. Academic members representing regions of Canada were to be identified by the co-chairs “in co-operation with members of the organizing committee – Hansen, ­Dombowsky and Yuel.” I, along with Alex McAuley from Victoria (and ­former chair of the NSERC peer review committee) and David Pink from Saint Francis Xavier (who had also been on the committee), were suggested as likely members. None of us were ever approached, to the best of our knowledge. The action plan also proposed the development of a “briefing book” in both official languages describing potential users of CLS, frequently asked questions, what Canada and industry spends abroad, other synchrotron sites, the current status of light source science, international commercialization activity, and the effort to “develop the Canadian argument.” This was to be assembled internally at the Ministry of Economic Development with contributions from SAL, and coordinated by Chris Stakiw, an administrative assistant to Brian Hansen at the Innovation Place office. A second document, the National Blueprint Document, would focus on the theme of enhancing Canada’s competitiveness by addressing governance and management structure, capital and operating finance options, and industrial * From the private files of Dennis Johnson.

The Canadian Institute for Synchrotron Radiation  67

and academic collaborations. The lead for this activity was to be Brian Hansen, with “consultative support from David Dombowsky.” The document would “basically be an assembly of materials arising out of existing m ­ aterials from the socio-economic impact assessment.” More will be said about the ­socio-economic impact study in subsequent paragraphs. The document was to be completed by the end of August 1996, a very short timeline. A “Monthly Bulletin Update” was to be created and circulated to key decision-makers in Ottawa and other strategic people, including Saskatchewan MLAs and MPs. Neither the Monthly Bulletin Update nor the National Blueprint Document ever materialized! The grand strategy outlined in the Hansen action plan was to have CLS Team Canada make presentations to “Tier 1” ministers in Ottawa, including Ralph Goodale (MP from Regina–Wascana), John Manley (industry minister), Jon Gerrard (MP from Portage–Interlake and secretary of state for science and technology), Lloyd Axworthy (MP from Winnipeg–Fort Garry), Anne ­McLellan (MP from Edmonton Northwest), Paul Martin (minister of finance), and Marcel Masse (MP from Frontenac). “Tier 2” briefings were to be held with Joyce Fairbairn (government leader in the Senate), David Dingwell (minister of health), Art Eggelton (minister of international trade), Sergio Marchi (­minister of the environment), and a host of other elected officials, committees, and ­bureaucrats. A CLS-sponsored reception and presentation was to be held in Ottawa with all Saskatchewan MPs invited. CLS Team Canada was also to meet with Dr. Arthur Carty, president of NRC; the purpose was to present the National Blueprint Document to NSERC and the Advanced Technology Association. A breakfast or luncheon with members of the NSERC Council was to be arranged in October, when the council was scheduled to visit Saskatoon to review the USask proposal. CLS Team Canada delegations were to visit third-generation synchrotron communities in Sweden (MAX-Lab), the APS near Chicago, and the ALS at Berkeley, the latter during an International Light Source Users Conference in October. The purpose of these site visits was to learn how ­third-generation facilities had, through advanced technology, “generated new economic ­ ­activity and commercialized new products and services.” Indeed, Hansen and ­Dombowsky visited several US facilities, to the chagrin of Skopik, Johnson, and the whole Canadian academic community that had worked so hard over the decades for the project. The cost of all these visits, the retainers for the Team Canada members, and the new office is not known, but the provincial expenses in this two-year period must have been well over $1 million. The creation of CLS Team Canada was to be announced as part of a ­national media and communications strategy. Among the fifteen implementation ­activities described in the action plan, a focus group program was to be ­delivered by “SAL and other light source scientists in Canada” to help coordinate a­ cademic/industrial collaborations. This was to be coordinated by the CLS ­office but managed by

68  The Canadian Light Source

Dennis Skopik. This activity had been discussed by the previous Kitchen Cabinet committee and was already underway without provincial involvement. The proposed cost of the construction and scheduling timelines were to be reviewed by Innovation Place staff, who were said to “have experience with major technology construction programs.” As a member of the Saskatchewan Opportunities Corporation’s Board of Directors, to which Innovation Place ­reported, Dennis Johnson was quite aware that Innovation Place had no experience in the construction of this type of scientific installation. Upon receiving the Brian Hansen action plan, Johnson sent President Ivany, Vice-President Tony Whitworth, Dennis Skopik, and Doug ­ Richardson a four-page memo containing his comments. He pointed out that the ­initiative to ­recruit CLS was SAL’s, an integral entity of USask; and that the u ­ niversity ­normally functions under an act of the provincial government and “at arm’s length.” In his opinion the action plan indicated that the province was ­embarking on a course of action independent of the university. CLS Team ­Canada was to be composed initially of six to seven industrial members and two to three representatives from academia, all of whom were to be determined by Hansen, Dombowsky, and Yuel. None of these three had any affiliation with the university. Johnson received no formal response, written or oral, from the senior university administration, perhaps because no one wanted to jeopardize the $10.5 million funding already committed by the province or to strain government relations in any way. After all, the province was at the time the biggest funder, and it was not a good idea to antagonize a critical funder. There was, however, a lot of “hallway talk.” Johnson did receive a letter from Murray McLaughlin, deputy minister of agriculture and a member of the original committee. McLaughlin’s letter ­suggested that the name “Team Canada” was a negative and could lead to confusion. He pointed out that the name had been used by the prime minister for federal missions abroad, and it also had been used by sporting teams. Johnson responded by saying that the name was introduced by Brian Hansen. The rationale given for forming CLS Team Canada was “to enhance Canada’s science and technology stature in the international community and enhance our competitiveness abroad by positioning the CLS project as an essential piece of Canada’s emerging technology and infrastructure strength.” This was, of course, what synchrotron users from academia, formally through CISR, had been telling elected officials and bureaucrats for years, and the claim had been an integral part of the strategy of both the USask and Western bids. The initial document mentions two co-chairs of Team Canada but did not elaborate on who were proposed as chairs. The formation of CLS Team Canada was discussed at a meeting of the new Executive Committee held in President Ivany’s office on September 3, 1996. Mr. Allan Taylor, a Saskatchewan-born former president and CEO of the Royal Bank of Canada, was proposed as

The Canadian Institute for Synchrotron Radiation  69

the ­industry co-chair and as spokesperson for Team Canada. There was considerable discussion about the time commitments expected of Team Canada members, particularly of Mr. Taylor. Brian Hansen seemed to think that senior executives from industry would spend considerable time and effort on behalf of Team Canada, and that a ten- to fifteen-day commitment from Mr. ­Taylor would be appropriate. Jim Yuel suggested that Mr. Taylor would likely not be interested in meeting with bureaucrats, and that three days would be more ­appropriate for a busy person. A draft briefing document was circulated from the new CLS office at Innovation Place in early September 1996. Johnson concluded that it was certainly not readable for the intended audience of industry leaders and senior bureaucrats. He recommended major editorial comments and corrections, as did others. The document described a synchrotron and partially outlined the chronology of events leading to the award of CLS to USask. It also provided some information on the use of synchrotron-based science, its potential benefit to Canada, the projected capital costs, and the existing financial commitments. It provided a table of the “Canadian Light Source Investment Requirements” projected over five years, totalling $115,856,000. Nearly all of this information was covered in both the earlier Western and USask proposals and the subsequent joint proposal. Led by the government representative, Brian Hansen, the Saskatchewan government and USask paid $39,000 for a socio-economic impact report for CLS, which was released by DRI Canada, a subsidiary of Standard and Poor’s. This report claimed that CLS could have a total economic spin-off of $35 million annually – a great contribution in the eyes of the Saskatchewan Department of Economic Development. But this report was recognized as unreasonable by the senior bureaucrats, such as NSERC president Tom Brzustowski, NRC president Arthur Carty, and other leaders in the Canadian scientific community. There were at least two problems with the $35 million estimate: no synchrotron in the world was generating annual industrial revenue of over $1 million, and no other synchrotron had generated estimates for their total economic spin-off. During this period, there were many meetings with CISR and SAL staff to further define the costs of the accelerators and beamlines. An additional meeting with the NSERC Council was held in Saskatoon on October 9, 1996, with a breakfast hosted by Premier Romanow. Dennis Skopik provided an overview of the CLS proposal and its relation to existing CLS infrastructure, and I gave a talk focusing on scientific and industrial opportunities at CLS. There was a tour of SAL, followed by discussions of how the combined proposal related to the McAuley recommendations about cost and beamlines. After lunch there were discussions of industrial strategies, project management, and operating budgets. The provincial representatives were not involved in this meeting.

70  The Canadian Light Source

In late October 1996, NSERC formally endorsed the CLS project and a­ dvanced it for funding consideration. This formal approval was very important, but again, where was the money coming from? NSERC had made it clear much earlier that they could not fund CLS. The most likely funder, Industry Canada, was still not proactively supportive. Moreover, the federal deficits continued, and it was not likely that CLS would be funded until these were tamed. So there was still no obvious funder (or owner or manager). Team Canada was seen by the province to be very important in getting Ottawa to decide how this facility would be funded and managed. In order to establish Team Canada, letters were sent in October 1996 to twenty-one people in industry and academia on behalf of Allan Taylor, who signed as “Special Advisor to the Canadian Light Source.” The letters asked the recipients if they would participate “with a team of Canadian Industrial and Academic leaders in directing Ottawa’s attention to this significant investment opportunity.” Most of the individuals contacted had either been visited by Doug Richardson, Dennis Skopik, and Dennis Johnson on trips to eastern Canada, or had written letters of support before the McAuley committee had endorsed the Saskatchewan proposal. The recipients were “to consider joining the CLS Team and committing to attend one Ottawa meeting with the Prime Minister and other senior Cabinet Ministers to advance the CLS project.” The nineteen-member Team Canada was then established. It included university presidents (for example, George Ivany from USask, Paul Davenport from Western, Rod Fraser from the University of Alberta, George Connell, formerly of the University of Toronto and Western, and Bernard Shapiro from McGill) and industry leaders (such as Allan Taylor, the Honourable Donald M ­ acdonald, C.E. Childers from Potash Corp, Doug Richardson from ­Mckercher Mckercher and Whitmore, and Bernard Michel, president of Cameco). Either Doug Richardson or Donald Macdonald (Liberal politician from 1962 to 1977 who had occupied such high cabinet positions as minister of finance under Pierre Trudeau) invited influential governing politicians mentioned earlier, Paul Martin, John Manley, Marcel Masse, and Jon Gerrard – who was born and raised in Saskatoon – to a meeting in Toronto on January 28, 1997, with representatives of the nineteen-member team. Macdonald said in this letter to Team Canada: “I support the submission which they will present to you for a capital commitment by the Government of Canada, and for ongoing support in operating costs. I note the support of NSERC for the CLS facility and its endorsement by an International ­Review Panel. Continuing effort in basic science is fundamental to the continuing ­success of Canadian high technology on world markets; and I respect the judgment of informed scientific opinion, both within Canada and internationally, that this facility is an essential element in future Canadian competitiveness.” Letters of support from eighteen Canadian universities and over forty

The Canadian Institute for Synchrotron Radiation  71

industries were attached, mostly as a result of Dennis Skopik, Dennis Johnson, and me criss-crossing the country in late 1996 and early 1997. CLS Team Canada travelled to Ottawa in late January 1997 and held meetings with John Manley and Paul Martin, the ministers of industry and finance, respectively. It is not clear how many attended that meeting. Neither Dennis Johnson nor I (and no one from the scientific community) were invited to the January 28, 1997, meeting. We were not annoyed because most of us thought it was embarrassing and potentially very damaging. We have no idea if the Team Canada effort had any influence on the outcome, which, as indicated below, must have been virtually decided by the time of the meeting. Team Canada may have increased the profile of the project in industrial and government circles, and it may have had an effect on the imminent formation of the CFI (which, unknown to us, was already decided on but not announced) to fund research projects such as CLS in Canada. We both think that the science lobby through NSERC, CISR, and MRC was much more effective. Following the Ottawa meetings, letters were forwarded from the CLS office to the members of Team Canada under President Ivany’s signature giving a report on the meetings. The report on the Manley/Martin meetings indicated that the project would be considered under a science and technology infrastructure program and “conveyed the view that no project was higher on the list” of science and technology project initiatives than CLS. The ministers had also conveyed the message that the days of single-source financing had ended, and that operating costs were a separate issue. Following the Team Canada visit to Ottawa, letters were sent from Ivany to Ministers Manley and Martin thanking them for meeting with Team Canada and for appointing Andrei Sulzenko and Paul-Henri Lapointe, deputy ministers of industry and finance, respectively, who were “to start discussions with Team Canada and the proponents of the project.” On February 18, 1997, letters were forwarded from Doug Richardson’s ­office to Sulzenko and Lapointe, as well as to Anne Park, acting deputy ­minister, ­science and technology, to confirm meetings in Ottawa on the following F ­ riday (­February 21, 1997) as a follow-up to Team Canada’s Ottawa visit. The l­etters stated that USask would be represented by “ Mr. Tony Whitworth, Chief ­Financial ­Officer, Mr. Dennis Johnson, Vice President of Research, Dr.  ­Dennis Skopik, Head of SAL, and on behalf of the province Brian Hansen, Assistant Deputy Minister, Economic Development, and Mr. Gordon Nystuen, A ­ ssistant Deputy Minister, Department of Finance.” Doug Richardson made these ­Ottawa appointments on behalf of the new Executive Committee (page 63). His success in this e­ ndeavour was ­undoubtedly enhanced by his deep knowledge of the Ottawa scene and his ­political connections. The meetings were labelled as “informational.” An agenda for the February 21 meeting with the assistant deputy ministers was prepared for the USask contingent. The agenda items included descriptions

72  The Canadian Light Source

of the project and the capital and operating costs that Skopik and J­ ohnson had projected. Brian Hansen presented on the “Parameters of a Project Agreement” and Tony Whitworth presented on the “Structure of Negotiations.” Adam Hitchcock, the senior synchrotron scientist from McMaster University, represented CISR and spoke on the uses of synchrotron light. There was a discussion on the possibility of an innovation fund, which had been hinted at by Ministers Manley and Martin, as well as by the media. The most memorable outcome of this meeting was a statement by one of the federal deputy ministers, who, while remarking that it sounded like a good project for Canada, wondered why it would be built in a remote location like Saskatoon. Some of the less-travelled Saskatchewan visitors nearly fell off their chairs! A similar comment had been made some fifty years earlier when the university was seeking funds to develop the cobalt source, a facility that was to have worldwide use for radiation cancer therapy (see chapter 2, section b). During the meeting, government representatives indicated that they had “no negotiating mandate concerning the CLS Project.” They also indicated that the project would be suitable for application to a new innovation foundation, to be announced in the imminent February budget speech. At this time, it should have been clear that the federal government was not going to fund a synchrotron at USask on a one-off basis. There were already strong suggestions at both the senior ministerial and the deputy ministerial levels that the announcement of a science infrastructure program was pending. However, it was apparent that Manley, Martin, and Gerrard were very supportive, and that a decision on a possible mechanism for funding the project had already been made.

5 The Creation of the Canada Foundation for Innovation

a) Announcement of the CFI as the Possible Funder, and the Foundation’s Importance As noted above, the very important federal politicians John Manley, Paul ­Martin, and Jon Gerrard had all been well informed about the project and had visited SAL in 1996 and earlier. But while they were certainly generally supportive, there was no specific agreement through which a government ­organization could fund this facility. Many of the probable funders, such as Industry Canada, had received considerable budget cuts in an effort to get the overall Canadian budget in balance. More industry support was necessary, but it was apparent that if money appeared, Paul Martin would support the project. In a note to Dennis Johnson dated February 7, 1997, Martin wrote: “Perhaps in the next budget (due in a few weeks), we will find a tool to move to the next stage. I  hope we can find capacity to do things like this – invest in Science and ­Technology.” This sounded very promising, as if plans had been made to have a central fund for science and technology projects in Canada. And indeed, soon after the ­February 20, 1997, budget, when it became very apparent that the deficit had been slain, the Government of Canada announced that it had created the Canada Foundation for Innovation (CFI) as a separate entity “to build ­Canada’s capacity to undertake world class research and technology for the benefit of C ­ anadians,” with an initial fund of $800 million. Jon G ­ errard was quoted in the Saskatoon Star Phoenix on February 27, 1997, as saying, “­Saskatoon’s proposed synchrotron is exactly the kind of project the new Canada Foundation for Innovation is looking to fund. I would be very ­optimistic quite frankly.” What an amazing result for SAL, the synchrotron users, and the large number of people who worked so hard over the years to know that there was a likely major federal funder for CLS! In my long telephone interview with Paul M ­ artin in 2010, it was apparent that this initial $800 million fund (to become $7 ­billion in 2018) was set up, at least in part, to provide a mechanism for

74  The Canadian Light Source

funding CLS. The whole Canadian academic community, and not just the CLS project alone, would benefit. The creation of the CFI was a remarkable idea, especially ­coming from the minister of finance, and in my opinion the academic community in Canada should have congratulated Paul Martin much more than it did. The CFI became a very important part of C ­ anadian universities’ efforts to attract o ­ utstanding academics, and it was crucial for the purchase of the very best equipment for university, college, and hospital labs (as we will see in more ­detail in chapter 7). Following the announcement of the CFI’s creation in the February 1997 budget, President Ivany wrote John Manley and Paul Martin* summarizing the discussions during the meeting of the Saskatchewan contingent with the Deputy Ministers on February 21. The main purpose of the letter was to ask if, “given the timing challenges and funding constraints previously mentioned,” as well as the fact that CLS had already passed peer review, the project could be considered “in advance of the establishment of the innovation foundation.” This request reflected the increasingly desperate need for finances to keep the SAL group intact. John Manley replied* that the foundation would operate as “the government’s vehicle for providing financial support for the modernization of existing research infrastructure and [the] development of new research infrastructure. Once the foundation becomes operational, it would consider an application for funding of CLS to this arm’s length agency.” In other words, federal funding for CLS was still close to two years off. b) The Struggle for Funding and Control: The Formation of the Collaborative Committee, and Possible Provincial Control The major source of funding – the new federal CFI – had now been identified, but there was still an enormous amount of work to be done to identify a large number of other funders and decide on the ownership and management structure of the facility. It was immediately clear that the CFI would only fund 40 per cent of the total cost of CLS, with the remainder coming from municipal, provincial, and other federal agencies, as well as universities and industries. This caused a tremendous increase in work and complexity. Throughout the spring of 1997, Brian Hansen and federal officials laboured, with the help of Doug Richardson, to create yet another committee, the Collaborative Committee, as proposed by Ministers Goodale, Gerrard, and Manley. Its mandate was described in a letter from the three ministers to President Ivany and Dwain Lingenfelter, the Saskatchewan minister of economic development. This * The following quotations come from the private correspondence between George Ivany and John Manley contained in Dennis Johnson’s files.

The Creation of the Canada Foundation for Innovation  75

committee, with representatives from both federal and provincial governments, was encouraged to explore the capital project and the operating facility, and to consider “essential steps and strategy for further advancing consideration of the project.” The letter further outlined the committee’s purpose as providing “a forum whereby the parties can arrive at a common understanding of the synchrotron project’s essential characteristics in both its development and operating phases, and the steps required to determine how the project can best be achieved and maintained.” This wording was strikingly similar to the Hansen action plan on the creation of Team Canada, which apparently was no longer active. The federal government’s lead representative was to be Maryantonett Flumian, assistant deputy minister at WED. The committee was to have representatives from the private sector, USask, and the provincial and federal governments, each of which would name a senior person to the committee. President Ivany responded by indicating that the university’s delegation would be headed by Mr. Hal Wyatt, chair of the USask Board of G ­ overnors; its members would include Dennis Skopik, Dennis Johnson, and Tony W ­ hitworth. Ivany made it clear that he expected one or more of these three to attend ­meetings with Hal Wyatt. The letter ended by saying that the university looked forward to an expeditious resolution of the issues so that the CFI could receive the CLS application in its first round of deliberations. The timing for this first round had not yet been determined. On May 12, 1997, President Ivany wrote to notify Dennis Johnson that the CLS Executive Committee established the previous summer had completed its responsibilities and would no longer need to meet because the Collaborative Committee would become operational in the near future. Dwain Lingenfelter, on behalf of the Saskatchewan government, conveyed the province’s willingness to work with the Collaborative Committee. He also indicated that Brian Hansen, as “Executive Project Leader, Department of ­Economic and Co-operative Development,” would be the province’s lead ­person. ­Lingenfelter’s letter also indicated that an effort was currently underway to recruit Jim McFarland, president of Husky Oil, to chair the committee. Later, when McFarland declined to serve, Bernard Michel, president of Cameco, was recruited instead. Hal Wyatt from the USask Board of ­Governors, Larry Spannier from the Saskatchewan Ministry of Economic and ­Co-operative Development, and Onno Kremers and Doug Maley from WED were all important members of the committee, and all played a crucial role in promoting the project. As noted above, the government representatives, headed initially by Hansen, wanted to control the industrial business generated by CLS. To this end, they suggested that there be two organizations to own and manage CLS: CLS Inc., which would run the accelerators, and Beamline Development Inc. (also ­referred to as Beam Inc.), which would operate the beamlines. They ­proposed to control industrial usage and generate revenue from industry using

76  The Canadian Light Source

25  per  cent of the overall CLS beamtime. This so-called two-headed monster was recognized by the scientists and many in the federal government as ­unworkable almost immediately, but it still took well over a year to get rid of it. For example, in May 1997, Dennis Skopik retained Bob McAlpine (the former NSERC administrator and physicist) to work for USask in connection with the Collaborative Committee. He produced a document entitled “The Governance Regime for the Canadian Light Source,” which did not include the “two-headed monster,” but this was ignored. Dennis Johnson wrote to Dennis Skopik on February 4 and 19, 1998, ­expressing his concern about the province’s concept of having Beam Inc. manage the industrial business. He also pointed out that (1) the three CSRF beamlines, which were owned by Western, would not be controlled by Beam Inc.; (2) there would be great difficulty in managing the overall facility if there were two separate entities managing the accelerators and the beamlines; (3) there would be a very negative response from the academic community (represented by CISR), which would be the major users of the facility; and (4) there would be inevitable problems with intellectual property ownership. Hansen seemed to have the idea that a single visit to an SR source would generate a marketable product and a revenue stream. This illustrated his lack of understanding of science and intellectual property ownership. Johnson pointed out that since ­USask had to be the applicant on the CFI proposal, there had to be problems with the CFI review if the province was operating the i­ndustrial business. He also ­emphasized that it was SR scientists that would build beamlines, and that it was not reasonable to have a provincial Crown corporation owning and managing them. Dick Batten, from Mckercher Mckercher and Whitmore, ­summarized a March 9, 1998, meeting with Hansen and Michael Corcoran (the new ­vice-president research at USask). He indicated that similar issues were discussed and that the difficulties had been pointed out to Hansen. The Collaborative Committee first met at Innovation Place in Saskatoon on July 29, 1997. Dennis Skopik, Dennis Johnson, Adam Hitchcock, Alex ­McAuley (chair of the 1995 NSERC site selection committee described above), and Clive Willis, formerly an NRC vice-president, made presentations. Willis was hired to produce a strategy and work plan. It was within this strategy that visits to Sweden and California were proposed so that the members could learn about synchrotrons. Johnson remembers that Dennis Skopik had to do a lot of work on the strategy document. In his review, Dennis pointed out that most of the work had been done already by the academic community and CISR. In May 1998, after having met for about ten months, the members of the Collaborative Committee were listed as Onno Kremers (federal government), John Wright (deputy minister of economic development in the Saskatchewan government), and Hal Wyatt (representing USask). Brian Hansen was listed as representing the CLS Secretariat. Dennis Skopik from SAL attended most meetings. Peter Wyant (provincial Department of Finance) frequently attended

The Creation of the Canada Foundation for Innovation  77

in a consulting capacity related to the projected finances and cash flow. The committee frequently met via telephone conference. Others, including Tony Whitworth, Michael Corcoran, and Dick Batten, attended some meetings on behalf of the university. Clive Willis attended as a consultant. Another new committee, sometimes labelled the “CLS Group” or the “USask Steering Committee,” was created by President Ivany in the summer of 1997 to advance the interests of USask. It did not include the provincial government group of Brian Hansen, Jim Yuel, or Jane Horachek! Throughout the summer and fall of 1997, the CLS Group (consisting of Ivany, Whitworth, Skopik, ­Corcoran, Richardson, and Wyatt) continued to meet. However, Brian Hansen was still active; indeed, he approached me in Ottawa a year later about becoming the director of CLS. Michael Corcoran replaced Dennis Johnson as associate vice-president in the summer of 1998, but Johnson attended as a consultant to the committee. The Collaborative Committee was active through the spring of 1998. It ­became the vehicle for securing a role for the proposed Beam Inc. in the ­funding plan and the governance structure of CLS, for reviewing and securing sufficient finances, and for exploring the role of the federal government in the governance structure. These meetings were in preparation for the USask application, which was to be submitted to the CFI by June 1, 1998. The presence of federal government representation on the Collaborative Committee, and the search for additional capital and operating funds for the project, resulted in the drafting of a contribution agreement between CLS Inc. and NRC. The president of NRC, Arthur Carty, had been very supportive of the project and had mandated Clive Willis to be heavily involved. Federal representatives pointed out, correctly, that the federal government had historically given NRC authority for ensuring the effective management of several large scientific facilities, such as TRIUMF in Vancouver. The operation of a national facility has international ramifications, and that includes the concept of reciprocity. Reciprocity in science is an important principle. For example, CSRF near Madison, Wisconsin did not have to pay rent for beamtime; likewise, ­Canadian scientists had used synchrotrons in several countries without charge. It was also noted that providing access to some foreign users might be restricted from time to time. When the application to the CFI was submitted on May 28, 1998, the secured contribution from the province of Saskatchewan was listed as $20 million, an increase of $10 million over the original commitment of $10 million made in 1996, prior to the peer review by the McAuley committee. Documents indicate the purpose for the increase in funding was to use the proposed new Beam Inc. Crown corporation to leverage funding from other entities for construction of beamlines. This was the first reason for the province’s quest for increased oversight of the project. A second reason might simply be called “control.”

78  The Canadian Light Source

The expected federal contribution (other than the CFI) was listed at $14 million; this was to come from WED, the Western Canadian arm of ­Industry Canada. It was reasonable, then, to have federal representation on the Board of Directors. Dick Batten in consultation with Louis Robayo, senior counsel for NRC, had drafted a unanimous shareholders agreement describing CLS Inc. as a non-profit entity, and suggested possible members of a CLS Board of ­Directors. In correspondence between Dick Batten and Bill Smith from the office of the NRC vice-president research about the above documents, there was a strong opinion that NRC should have an expanded role on the CLS Board, and that Arthur Carty, president of NRC, be the initial chair of the board. This was an excellent choice because Arthur was an incredibly effective administrator and scientist. Although the reasons are not stated, Smith had concerns about the CLS Steering Committee and wanted it eliminated; in its place, he advocated the formation of two boards: one for CLS Inc. and one for Beam Inc. However, NRC favoured a unitary structure. So, the key question was still: Who would then apply for the CFI funding and manage the construction until the facility was operating? At this stage provincial representation at various meetings indicated greater provincial oversight, first with the involvement of John Wright, president of the Crown Investments Corporation, and then of Fraser Nicholson, deputy minister of economic development. In July 1998, Larry Spannier, an assistant deputy minister from the Department of Finance, replaced Brian Hansen as the provincial government’s representative on the project. Doug Maley of WED also joined the committee. Both proved to be highly regarded by the university and CISR, and both were effective proponents of the project for several years. During this period, considerable effort was directed toward ensuring the ­accuracy of the capital and operating cost estimates based on initial cost ­analysis by Dennis Skopik and the SAL staff working with UMA and CISR representatives. Peter Wyant from the provincial Department of Finance, working with Dennis Skopik and Clive Willis, reported to the Collaborative Committee. Agra-Monenco was hired to perform a review of costing and contingencies. By April 1998, draft legal documents describing agreements between CLS and a “Crown Corp.” (Beam Inc.) had been prepared. Another draft agreement between USask and the province would allow “the Government of Saskatchewan to exploit the research capabilities and the industrial potential of the facility by establishing partnerships and associations with national and international users for the establishment, construction and funding of the ­operation of beamlines and experimental end stations.” The agreement also ­indicated that Beam Inc. would “assume primary responsibility for the design, construction and utilization of the beamlines and work research stations so as to maximize the economic impact of the facility and the responsibility for the marketing

The Creation of the Canada Foundation for Innovation  79

of the facility to potential industrial and commercial users.” The proposed activities of Beam Inc. were described as follows: “First, to market the use of synchrotron light to industry, scientific institutions and governments; second, to identify ventures that would allow Beam Inc. to establish vehicles through which investments were to be made in the design, construction, ownership and operation of beam lines and experimental chambers for use at CLS; and third, to contribute to the ongoing support of the human and capital infrastructure in Canada associated with synchrotron light source at CLS.” The Collaborative Committee thus appeared to be taking the initiative of beamline development and industrial involvement, moving it away from the purview of the university proponents and CISR. There was no mention of the previous involvement of CISR. The fact that it had been agreed that all three beamlines at the CSRF facility in Madison would be repatriated to CLS seemed to be ignored initially, although this had been agreed to by the scientists and Western officials. However, in a draft memorandum of agreement between CLS Inc. and Beam Inc., it was later recognized that Beam Inc. could not own the Madison beamlines. When the concept of a Beam Inc. was first introduced by Brian Hansen in 1996 in meetings of the Kirk Hall–based Steering Committee (the Kitchen ­Cabinet), it certainly appeared that the government proponents thought that it was an opportunity to profit through revenues from gaining intellectual property or from rental income for use of the beamlines. This idea was informed by the propensity of successive socialist governments (CCF-NDP) to create Crown corporations, some successful, others doomed to dismal failure. As the project planning progressed into 1998, with the provincial government wanting to control beamline establishment and use in return for its investment, a debate b ­ egan concerning the amount of beamtime that would be allocated for use by the Canadian academic and scientific community. A “CLS Policy Note” from the provincial government stated that, while Beam Inc. was to recruit, install, and market the beamlines, ownership would rest with CLS Inc. This same note indicated that 35 per cent of the beamtime would be dedicated to the CLS staff and academic and scientific community after a peer review of applications for such use. Thus Beam Inc. was to control 65 per cent of the facility’s available beamtime! This was later negotiated down to 50 per cent, as indicated in the draft memorandum. It appeared that profit was the province’s overriding interest. In retrospect, the concept of a Beam Inc. did result in a $10 million increase in provincial funding. As we will see below, Beam Inc. was finally discarded during the final negotiations with the CFI. The CFI insisted on a “unitary” governance and operational structure – which was a relief for us. Perhaps the most ironic aspect of these machinations was that neither the provincial nor the federal representatives on the Collaborative Committee were

80  The Canadian Light Source

scientists, much less scientists who used synchrotron radiation. Dennis Skopik, who participated in all Collaborative Committee meetings, however, had a strong science background, in addition to his senior administrative experience at the advanced scientific installation SAL. As might be expected, Dennis was very frustrated by many of the provincial government’s activities. The final meeting of the Collaborative Committee was held on June 29, 1998. In July, Bernard Michel (in a letter to USask president George Ivany, the federal ministers John Manley, Ralph Goodale, and Ron Duhamel, and Janice ­MacKinnon, the provincial minister of economic development and co-operative development) stated that the committee had in large part fulfilled its mandate and should be either disbanded or re-established with a different mandate. In either case, he stated, the current committee would be disbanded upon the announcement of the CFI’s decision. The long letter described the origin of the committee, its mandate and accomplishments, and listed activities needing attention, such as the design of the facility and the finalizing of the governance structure that would oversee CLS. Some of the listed activities had been established before the advent of the Collaborative Committee. After reading the letter, USask people felt that the Collaborative Committee or its replacement went well beyond its original mandate. Tony Whitworth wrote that the university administration thought the letter was crafted by Brian Hansen. While these negotiations were going on, the expanded USask ­Steering ­Committee continued to meet with the addition of Hal Wyatt (board chair, USask), Larry Spannier (Saskatchewan Department of Economic and ­ ­Co-operative Development), Doug Maley (WED), and Peter Hackett (NRC). Fraser Nicholson, the Saskatchewan deputy minister of economic and ­co-operative development, attended when available. An internal university project steering committee was established for the purpose of assuring that the CLS project team fulfilled the financial, administration, and reporting ­requirements of the funding partners. Chaired by Laura Kennedy, the committee had representation from CLS, UMA, provincial and federal funding agencies, and appropriate staff from the USask financial/purchasing and the facilities ­management groups. c) The CFI Application, Financial Problems, and Industrial Workshops to Build the Industrial Profile In addition to the ownership and control issues outlined above, there were still great difficulties when it came to raising all the capital and operating funds ­required. The CFI would only pay 40 per cent of the total costs of any project, and a significant part of the 60 per cent matching money had not yet been identified. In addition, the CFI would not provide any of the projected $8.6 million annual operating cost in 2004. This turned out to be an unrealistically small estimate.

The Creation of the Canada Foundation for Innovation  81

On June 25, 1997, the first $800 million CFI innovation competition was announced in Ottawa. The deadline for preliminary applications was June 1, 1998; once these proposals were vetted, those invited to reply were given until October 1, 1998, to submit their final applications. In addition to the equipment funds mentioned above, the Canada Research Chair (CRC) program was ­announced in December 2000, with the goal of recruiting outstanding ­researchers and providing needed research infrastructure for hundreds of new university faculty across Canada. As an aside, Dennis Johnson was planning to step down as associate vicepresident research at USask. I applied to be his replacement, because I thought I could really help USask get CLS established and ensure that it would have a positive effect on a large part of the campus. I applied despite the very negative friendly advice of Dennis Johnson and Dennis Skopik, who thought that the politics of the university dictated that a CLS supporter like me would never get the job. Dennis Johnson had spent a good deal of time on this project in the last few years, and there was little chance that another CLS supporter would be hired who would spend enough time raising money for the non-synchrotron part of the university faculty. I nonetheless interviewed for the job in early 1997, before a thyroid operation on February 20, 1997. In the end, I was ­indeed unsuccessful. The very well-qualified committee had strong research representation from across the campus, especially in science and medicine. The successful candidate, Michael Corcoran from the University of Victoria, started in August 1997. I had broad interests – including many aspects of music, religion, and sports – in addition to science, but the committee was not interested in hearing about those interests at my interview, or my suggestions for helping the music, religion, and humanities areas at USask, as well as CLS. I believe that the potential financial liability from CLS (still unfunded at that time, with a potential large financial risk to the university) masked rational discussion and decision-making. Fortunately, this negative decision did not reduce my enthusiasm for, or commitment to, the project. In June 1997, the rules governing the funding applications to the CFI had not yet been published. Fortunately for CLS, a separate regional/national ­facility category was created by the CFI, a decision from which CLS benefited greatly, both initially and over the next twenty years. There were very few national ­applications of any kind in the next fifteen years, and certainly only one or two others of such a high quality. As a result, CLS was successful with nearly every application made in the course of the next decade, and it received many times the initial $56 million from the CFI. Obviously, the CFI was interested in funding CLS, and the science had been approved in the aforementioned NSERC decision to go ahead with the project in June 1996. However, NRC and NSERC in Ottawa encouraged CISR to hold a number of regional and national workshops in early 1998 to “increase the

82  The Canadian Light Source

awareness of the uses of SR mainly in major Canadian industrial ­sectors” so that Industry Canada could better support CLS. It was critical that industry ­become a strong user of the facility; and indeed, the initial CLS policies (­ initiated by the province of Saskatchewan and supported by CISR) dictated that 25  per  cent of the user time could go to industry if required. Clive W ­ illis, formerly ­vice-president of NRC, encouraged by the NRC and NSERC ­presidents, was immensely helpful in getting these five workshops organized in a very short time period. They were devoted to biotechnology, biopharmaceuticals, and medicine; mining, natural resources, and the environment; ­materials and manufacturing; telecommunications; and a general workshop. Four of these were held in Toronto, Vancouver, Hamilton, and Ottawa in the five weeks ­between December 31, 1997, and February 4, 1998. Adam Hitchcock and I were mainly responsible for phoning the many potential speakers and attendees, while a few dedicated synchrotron users helped organize each workshop; industries and government organizations supported the workshops financially. Clive W ­ illis did most of the work booking the hotels, setting up schedules, and writing up the final reports. Attendance at these workshops ranged from ­thirty-one in Vancouver to fifty-four in Ottawa. The industrial participation at all workshops was over 40 per cent, with the remainder of attendees coming from academia and government. Detailed reports on all workshops were available. These workshops provided many new academic and industrial members for CISR, many new ideas for research at CLS, and resulted in some later industrial users at CLS after 2005. Dennis Skopik was the primary author of both the initial June 1, 1998, submission to the CFI and the final October 1, 1998 application. To raise the initial federal matching funds for the June 1 submission, Doug Richardson ­organized a meeting (held on May 21, 1998) at the Saskatoon airport with Paul Martin, Dennis Skopik, Dennis Johnson, and Bernard Michel, chair of the ­Collaborative Committee. Michel made the “presentation” asking for ­additional funds from the federal government. This came through within two weeks. This was pivotal, and George Ivany was able to present the June submission to CFI president David Strangway. The science and the project had already been accepted by the NSERC committee in 1996, so the only problems had to do with the budget, ownership, and the management structure! The request to the CFI was for $71 million out of a total budget of $178.2 million (the $71 million is 40 per cent of $178.2 million as dictated by the CFI). A number of the proponents went to Ottawa for the CFI Board meeting on October 6, 1998, including a substantial delegation from the ­Saskatchewan Ministry of Economic Development (Fraser Nicholson, Larry Spannier, and Brian Hansen), followed by other meetings in Saskatoon on ­October 15 and 16, 1998, and again in Ottawa on October 26, 1998. At this time, the provincial ­representatives were still travelling around independently

The Creation of the Canada Foundation for Innovation  83

of the university group to develop the industrial business without any expertise to do so. O ­ bviously, it was still not known who was going to own, manage, and control CLS, and damaging conflicts were evident between the university group and the provincial group – conflicts that were very frustrating to Dennis Skopik and all of President Ivany’s group. I had several helpful phone conversations with Larry Spannier toward the end of 1998, and there was a general feeling that the province wanted to hire me to help with the project – but hired by Beam Inc. or CLS Inc.? Indeed, it was Hansen who first offered the director’s job to me in Ottawa on October 26, 1998. Things were getting complicated. The CFI was quick to reply to Ivany in late 1998. The letter stated that the large $22.7 million in-kind contribution from the accelerator plus land at SAL would not be allowed as matching funding, and so the total budget became $140.9 million. Matching money became quite a “game” in the CFI applications across Canada, because it was so difficult in most cases to generate the specified 60 per cent matching funds. Because of the 40 per cent matching rule, this meant that the project became a $140.9 million project with a request to the CFI of $56.4 million. This led to great difficulties, because more contributions were required to come close to the $84.5 million matching funds. When Ivany’s Steering Committee then scrambled back to the province, the city, and WED, each increased their commitment. Again, Doug Richardson played an important role in these negotiations. The matching funds now stood at $25 million from the province, $2.4 million from the city, $21.8 million from the Government of Canada (namely WED), $7.3 million from USask (mostly savings from the SAL power bill of $1 million per year during the 1998–2005 shutdown), $600,000 from Western and the University of Alberta, $2 million from SaskPower for the utilities bill, and $6.5 million from NRC and Natural ­Resources Canada (NRCan). This totalled $65.6 million (see table 1). At that time, the $25 million from the province was to fund personnel and i­ nfrastructure associated with the proposed creation of Beam Inc. (Note how important WED was for the funding of this project.) But, $18.9 million in capital (see table 1) was still to be raised, and there were still no commitments for the estimated ­annual operating costs of $8.4 million for CLS Inc. and $5.5 million for Beam Inc. That $18.9 million shortfall could have caused a very large problem for U ­ Sask. The two most likely options were for USask to either pay the $18.9 million or withdraw from the project as there was still no legal commitment. A pivotal meeting was convened on February 8, 1999, to prepare for the CFI review on February 22, 1999. The latter meeting was not to be a scientific review (the project had already had a successful NSERC scientific review in 1996), but rather was concerned with reviewing the funding (capital and operating), organization structures, and management and planning issues. The following people were present at the February 8 meeting: Dennis Skopik and

84  The Canadian Light Source Table 1.  Summary of the Initial CLS Project Capital Funding ($M) Source of Funding

Total ($M)

Canada Foundation for Innovation (CFI) Province of Saskatchewan Western Economic Diversification (WED) University of Saskatchewan City of Saskatoon Universities of Alberta and Western Ontario National Research Council (NRC) Natural Resources Canada SaskPower Government of Ontario Government of Alberta Boehringer Ingelheim (Canada) Ltd.   Total Funding

56.4 25.0 21.8 7.3 2.4 0.6 4.5 2.0 2.0 9.4 9.2 0.3 140.9

Note: Every one of these thirteen contributions required a large written “agreement” of between ten and twenty pages. As noted in the text, for the University of Alberta and Western, there were two or three “extra” agreements, which in the case of Western totalled over two hundred pages.

Emil Hallin from SAL; Michael Corcoran (vice-president research); Dennis Johnson; Matt Webster from the office of the USask vice-president administration; Barry Hawkins (UMA); Mike Bancroft (Western); Ron Cavell (University of Alberta); Walter Davidson (NRC); Clive Willis (Collaborative Committee); Doug Maley and Ed Wiens (WED); and Peter Wyant and Larry Spannier from the province of Saskatchewan. Some of the provincial costing was still less than satisfactory, the industrial revenue was overestimated, and the structure with the “two-headed monster” was not reasonable in our estimation. The high-level external evaluation committee meeting, held on February 22, 1999, and organized by the CFI, met with the same CLS representatives listed above. As mentioned, the CFI committee had the mandate to review the proposed funding (capital and operating), organization structures, and management and planning issues. Catherine Armour, coordinator of institutional relations, and Denis Gagnon, the senior executive advisor, represented the CFI. ­Catherine organized the meeting and acted as secretary. The committee had strong representation from the engineering and business sectors as well as high-level synchrotron managers from international synchrotrons (for example, Bill Cochrane, president of W.A. Cochrane and Associates, acted as chair). Colin Franklin from Nepean, Ontario, and Francis Hartman from the Department of Civil Engineering at the University of Calgary also represented the engineering and business sectors. Three very senior international synchrotron administrators were the synchrotron reviewers: Jean-Louis Laclare, director of Project Soleil in France, Peter Lindley, director of research at the European

The Creation of the Canada Foundation for Innovation  85

Synchrotron Radiation Facility (ESRF) in Grenoble, and Stephen Milton, synchrotron ring manager at the APS in Chicago. The CFI committee wisely insisted that a new company owned by USask, CLS Inc., should build and operate the entire new facility. CLS Inc. would have its own independent board reporting to the university, with no Beam Inc. ­involvement. It was recognized that the two-headed monster would not work. The provincial government had lost its attempt to manage part of the facility, but fortunately remained as a principal funder. In addition to the amounts given above for capital, NRC and NSERC (in response to the CFI committee wanting further commitments) promised operating funds, both at the meeting and shortly afterwards. Arthur Carty (president of NRC) and Tom Brzustowski (president of NSERC) had each been enormously supportive, and they obviously had worked together. Even at that stage, without their strong support, the project could have been derailed. The CFI Multidisciplinary Assessment Committee (MAC) met in Ottawa to review the “first round” proposals from March 8 to 10, 1999; this was shortly after the CFI Site Review Committee submitted its report. Both the Site Review Committee and the CFI MAC recommended funding for CLS; and on March 31 the CFI announced full funding as requested. The press release from Saskatoon said that “CLS represents an unprecedented level of collaboration of governments, universities and industry in Canada.” Eighteen universities, in addition to USask, endorsed the CLS project on behalf of over three hundred users of synchrotron light in Canada. (For more, see my letter to all CISR users in appendix 2.) As per the CFI award agreement, as well as George Ivany’s letter to the CFI, the university agreed and committed to the following: 1 Develop a Facility owned and controlled by the University of Saskatchewan that will include an accelerator, booster, storage ring and at least six third generation beamlines by the time of the commissioning date of the Facility. 2 Secure sufficient capital funds to complete the construction and commissioning of the facility while not requesting further funding from CFI for the capital costs of the Facility. 3 Ensuring that there will be an annual operating budget of $13.8M ­committed, upon commissioning of the Facility, expected in 2003. 4 Owning and controlling for a minimum of five years from the date of ­commissioning of the Facility, the non-profit organization established to provide an effective and unified management structure known as the ­Canadian Light Source Inc. This announcement was an enormous “high” for everyone associated with the project. The chance of getting this facility accepted and funded in 1991 was

86  The Canadian Light Source

very small. The number of stars that had to fall in line (as described above) was enormous, and the chances of them lining up were very small. But enormous effort from a large number of people working mostly together, in addition to a good deal of amazing timing and a lot of luck, enabled our success. In hindsight, if the CFI (or equivalent) had not been formed for another year or two, a Canadian synchrotron facility would probably have been funded within five years, but likely not at USask or Western. The SAL facility would have closed down, and most of its employees would have gone in various directions. Without the talent and dedication of Dennis Skopik and his SAL colleagues, and with so few synchrotron users at USask, there was little chance that USask would have been able to mount a later application to the CFI. The project could not have ended up at Western, even with a lot of synchrotron users. With Paul Davenport as president until 2009, it was highly unlikely that Western would have got behind the project. This meant that it would have likely gone to NRC in Ottawa, or to Toronto or Vancouver. Even with the funding secured, Dennis Skopik left for a big nuclear physics laboratory in the United States. It was very surprising to everyone when Skopik announced that he was leaving, after all his efforts. But he was an American citizen who came to Saskatoon in 1970. He returned to a senior administrative position at a very well-known nuclear physics lab, Thomas Jefferson National Accelerator Facility in Virginia. Dennis had told me that if I came to CLS as director, he would remain as machine director. Dennis knew that it was important to have a director who was well known in the SR area, and so he suggested to the university that I assume this role. I certainly looked forward to working with him after so many positive and productive interactions in the 1990s. Why did Dennis leave? As a dedicated scientist, his first priority was to continue in the field to which he had devoted himself for over forty years – nuclear physics. Also, his aging mother-in-law lived in Washington, very close to the Jefferson facility in Virginia. He was undoubtedly worn down by the enormity of the job of getting the facility accepted and mostly funded and the often negative provincial and university politics. Some of the above reasons were also important factors in my decision to step down as interim director of CLS two years later.

6  My Role as Interim Director, 1999–2001

a) Introduction In early April 1999, I was approached by Tony Whitworth, vice-president administration at USask, to go to Saskatoon as interim director for a two-year term lasting from September 1999 to September 2001, seconded from Western. A new, very supportive dean of science at Western, Fred Longstaffe, paid my existing salary, and the salary of T.K. Sham, for two years out of his budget. Longstaffe’s generous support was very unusual. He probably hoped that ­Western researchers would benefit greatly from CLS even if it was located at USask. Overall, I was leaving a very positive atmosphere at Western, one characterized by many very friendly long-time colleagues and a very active research group. Fortunately, my senior colleague Masoud Kasrai supervised most of my graduate students in my absence. I only returned to London for a few days in the next two years because I was so tied up in Saskatoon. Both my wife and I looked forward to the new challenge and the friendly Prairie people. Joan and I drove from London to Toronto and then on to Saskatoon in late August knowing that my only sister Jane, an unmarried French professor at Scarborough College, would die in the next few weeks in Toronto. After a short and pleasant holiday at Lake Waskesiu, we moved into a nice rented bungalow in Saskatoon. After Jane’s death in mid-September, Joan and I returned to Toronto for her funeral and to settle her affairs. I made good use of this hectic nine-day stay in Toronto by setting up a meeting within a hundred yards of my sister’s apartment with David Bogart, executive director of the Ontario Innovation Trust (OIT). It was my hope that OIT would provide $9.4 million of the remaining $18.9 million that we needed to raise. I was surprised and delighted to find that Bogart was quite keen to help fund part of the shortfall at CLS, mainly because most of the present Canadian users of synchrotron facilities came from Ontario. Every province established an OIT-like organization to help with the matching money for CFI grants – just what the federal

88  The Canadian Light Source

authorities had hoped for. It would still be very unusual for Ontario to fund a facility in Saskatchewan. I returned to Saskatoon on September 23, 1999, and the sod-turning ceremony for CLS was held on September 27, outside on a cold and windy fall day. Many people that had played an important role in landing CLS were there, and they all needed sweaters and warm jackets to cope with the weather. My eighty-nine-year-old mother in Winnipeg was not well enough to come, but my ninety-five-year-old aunt Eileen Bancroft came from Regina with her neighbour, Gloria Erickson. They were both delighted to be there. I gave a talk on CLS at the Saskatoon Public Library that night. The local community was very excited about the facility, and I soon found that I was a bit of a scientific “rock star” in the community. For example, all the taxi drivers knew about the facility, and many local people visited the construction site in my two-year tenure at CLS. I had no job description for the interim director’s position when I arrived in Saskatoon (and indeed never did have one), but it seemed to me that I had five major tasks: (1) raising the $18.9 million shortfall; (2) getting an agreement between CLS and its owner, USask; (3) setting up the basic organization within CLS, the committee structure, and principles of operation; (4) planning for the initial beamlines (called phase 1 beamlines); and (5) encouraging USask to hire key synchrotron users at USask. As seen below, I had good success in some of these tasks, and partial success with others. b)  Raising the $18.9 Million Shortfall I arrived in Saskatoon believing my most important job was to secure the $18.9 million that the project still required (see table 1). In fact, the potential shortfall was much worse than $18.9 million. Because of the 60/40 CFI matching formula, the $18.9 million carried with it $12.6 million in matching funds from the CFI that could not be accessed until the first figure was raised. Thus $31.5 million was still required to complete the project, and this was most of the money needed to fund the beamlines and experimental equipment for users. It is important to note here that the $140.9 million initial funding included the operating budget for the first five years until operation (1999–2004), ­including all salaries. Fortunately, raising the operating budget was not an ­onerous additional task during my tenure. However, it certainly became a very difficult task every four or five years beginning in 2004. This fundraising from multiple sources created great difficulties for future CLS directors – namely Mark de Jong, Bill Thomlinson, Josef Hormes, and Rob Lamb – as well as ­Peter ­MacKinnon, the USask president until 2012, and Beryl Lepage, CLS’s chief financial officer from 1999 to 2018. But did CLS have the lead role in the initial fundraising, or did the owner, USask, have that role? Because the USask administration had no first-hand

My Role as Interim Director, 1999–2001  89

expertise with synchrotron facilities, it seemed obvious to me that I should lead the fundraising. But, in a few months it became apparent that part of the administration, their lack of synchrotron or research-funding expertise notwithstanding, thought that the university should be the lead fundraiser. Could there be a perpetual conflict between CLS and USask analogous to the one mentioned above between the province and CLS? Without the $18.9 million, there would be a building and accelerators but no users. And how many times had other provinces funded facilities in Saskatchewan? The initial odds of getting this money from Ontario and Alberta were not high. But I had absolutely no thought of saying to USask and the province that they needed to raise the $18.9 million, as this might have made it very difficult for the university administration, faculty, and staff. Because of my previous uniform success in raising money for many large projects at Western (e.g., Surface Science Western and CSRF in Madison), I was very confident that the remaining $18.9 million could be raised, as I stated in an initial meeting with the USask Board of Governors in October 1999. I think that very few, if any, of the faculty at USask believed me; and it was certainly possible that the money would not come from two (or more) provinces, leaving USask having to come up with the funds. Before negotiating with the provinces for financial contributions, it was necessary to set up some of the principles to clarify the general access rights that a contributor might expect for both financial and human capital contributions. Adam Hitchcock had drafted such a document in 1999. By the end of that year, Beryl Lepage and I prepared a ten-page discussion paper for the first seven beamlines, entitled “Beamline Issues: Ownership, Priority Access and Industrial Issues.” This later became the “Canadian Light Source: Framework for Contributions.” For example, we stipulated in these documents that the seven initial beamlines would be “national facility” beamlines, owned by USask and operated by CLS on behalf of the owner. Thus, a contribution from the provinces would not result in that province owning any of the equipment; and moreover, a contributing province would not hire people to operate any of the CLS equipment. These stipulations were similar to those at other international synchrotron facilities. There was little disagreement with the initial ­Ontario and Alberta provincial contributors on the ownership/operation issues. Also, the funding province would not receive priority access for its contribution; but the beamline developers from that province would receive priority access, without the normal peer review. In return for its financial contributions, a province would be offered a seat on the CLS Board. The amounts of priority access were sometimes very contentious, as we will see below. Within a month of meeting David Bogart of OIT in September 1999, I wrote and submitted an application to him for the $9.4 million (approximately half

90  The Canadian Light Source

of the $18.9 million shortfall). This application was shared with all the USask committees for information and comment. There were no adverse reactions. I also began to work with two chemistry colleagues at the University of A ­ lberta, Ron Cavell and Ron Kratchovil, to obtain a similar contribution from Alberta. Cavell was very important in promoting CLS in that province, and ­Kratchovil initially represented the provincial government to discuss the request for $9.2 million, leaving $300,000 outstanding (see table 1). The reaction from OIT to my first proposal was negative; I was asked to provide a business case for Ontario investing in CLS. On January 19, 2000, I submitted a 20-page proposal, with over 150 pages of supplementary appendices of information and lists of the many users from Ontario. Again I shared this application with the USask administration and all the committees, with no negative comments. Fortunately, Dr. Cal Stiller, a well-known Western medical academic born in Saskatchewan, was on the OIT board, and I believe he was very supportive of having the facility in his home province. Indeed, Dennis Johnson had served on the MRC Council with Dr. Stiller, and had invited him to speak in Saskatoon a few years earlier. A number of negotiations took place in Toronto and London. I represented CLS for the first meeting, and Beryl Lepage and Matt Webster (from the USask administration) joined me for later meetings. Beryl, seconded from the university administration, was a very important and positive player in these complex and difficult negotiations. For example, OIT initially insisted that if they gave us $9.4 million, CLS would have to guarantee that it would spend at least $9.4 million in Ontario. Although this might have been possible, there was no way that CLS could have an open bidding system and guarantee that $9.4 million would be spent in Ontario. We insisted that CLS could not agree to such a demand. Eventually seeing that such a stipulation was neither reasonable nor important, OIT backed down, as they did on other stipulations. (In the end, more than the $9.4 million was spent in Ontario.) The $9.4 million was awarded with no significant headaches for CLS or USask. By late 2000, I was quite confident that this money was “in the bag,” but the final, very large contract was not signed until July 2001. The money was to be used mainly to upgrade or replace the two soft x-ray beamlines from the Madison CSRF facility, with T.K. Sham as beamline leader. None of the OSC money was to go to any users in Ontario, such as the two most dedicated synchrotron scientists, Adam Hitchcock and T.K. Sham. Also, as noted above, Beryl and I insisted that provinces could not own any of the equipment at the facility that they were funding. This one agreement had many complex components. The Ontario contribution would be overseen by a new organization set up by the Ontario government, the Ontario Synchrotron Consortium (OSC), to be managed by Western. Setting this up required a lot of time from me and the legal manager from Western, Peter Ross, in 2001 and 2002. Because I was now back at Western, the loser

My Role as Interim Director, 1999–2001  91

of the competition for CLS was now the biggest CLS supporter! A 150-page document on the OSC from February 2002 included three large agreements: (1) the Ontario Synchrotron Consortium Agreement; (2) the Ontario Innovation Trust Agreement; and (3) the agreement between Western and USask. There was also a memorandum of understanding between the OSC and CLS. In the first agreement, five universities (Western, McMaster, Waterloo, Toronto, and Queen’s) agreed to form the OSC to oversee and manage OIT’s $9.4 million investment in CLS. Later, the Universities of Ottawa and Guelph also joined. Regular meetings were held to review the spending, and later to review the usage of the beamlines by Ontario scientists. Because of further investment from Ontario in CLS over the last fifteen years, the OSC has continued to meet annually. As the lead university of the OSC, Western signed the second agreement with OIT to receive, oversee, and manage the $9.4 million contribution. And then in the third agreement, Western, on behalf of the OSC, signed an agreement with USask to ensure that CLS spent the $9.4 million from Western appropriately. These contracts, agreements, and memoranda were all written in small, single-spaced “legalese.” The agreement with Alberta was, if anything, more complex, and I did not control the application first-hand as I did the Ontario application. The main Alberta-USask contract was very similar to the Ontario contract; but from my point of view, it was less satisfactory. The University of Alberta signed an agreement with USask in August 2001 to contribute $9.2 million on behalf of the Alberta Synchrotron Institute, which had just been set up. But there were complications. First, Matt Webster from the USask vice-president administration office became involved; and at times, Beryl and I from CLS were negotiating one issue while Matt was renegotiating the same issue without our knowledge. Who was responsible for these contracts, CLS or USask? Again, accountants at USask did not have the scientific or fundraising expertise to lead any fundraising at that time. Second, Doug Maley, from WED and based in Edmonton, had joined the President’s Committee; and like Larry Spannier from the Saskatchewan government, Doug was very supportive of the project and very helpful. WED often contributed financially to the project (see table 1 for WED’s very large initial contribution). In this case, WED hired a senior government official, Ken Alexei, to write the proposal for Alberta. Ken did a fine job, but the province would only give the $9.2 million if WED and the Saskatchewan government gave Alberta $5 million ($1 million a year for five years) for a new Alberta Synchrotron Institute at the University of Alberta. This money was to be used to help fund Alberta scientists to begin their synchrotron research at foreign facilities from 2001 to 2005, when CLS started operating. This should have resulted in a number of dedicated CLS users from the Alberta universities. Unfortunately, the money did not build the number of Alberta users significantly, and very few papers were published

92  The Canadian Light Source

by Albertans from work at US facilities between 2001 and 2006. The question remains: What was the $5 million used for? There are relatively few Alberta scientists using CLS even in 2018 (less than 7 per cent of the total users), considering the proximity of the large Alberta universities to CLS. As an aside, the Saskatchewan government and WED established the Saskatchewan Synchrotron Institute (SSI), directed by Dennis Johnson from 2002 to 2004, with $1 million in order to increase the use of SR by scientists, post-docs, and graduate students from the province. By June 2004, over seventy individuals in the Saskatchewan research and development community had used SR in their research programs as a result of SSI travel grants to foreign facilities. It is unfortunate that Alberta did not use its $5 million in a similar way. But this is just one example where a large amount of money (the $5 million to Alberta) generates very little effect, while a much smaller amount of money (the $1 million to Saskatchewan) generates a much greater one. Meanwhile, other time-consuming fundraising efforts in Manitoba (by me) and Quebec (by Rob Slinger, a CLS employee) were not successful. The University of Manitoba’s senior administration was not helpful. I employed Gary Filmon, the former premier of Manitoba, to help with the Manitoba effort and he did a fine job, but no money was contributed by the Manitoba government. I knew Gary Filmon’s wife Janice (now the lieutenant governor of Manitoba) well from driving her in a carpool to the University of Manitoba for two years in my third and fourth undergraduate years in the early 1960s. Rob had expensive Quebec lawyers involved with his effort. There was still $300,000 to be raised (see table 1), and I was promised that amount by Paul Anderson, a former Western student who was now vicepresident at Boeringer-Ingelheim, a large multinational drug company with a head office in Switzerland. As with most drug companies around the globe, they hoped to use SR in the future for determining the structures of proteins and protein-drug combinations to enable more efficient drug development. Amazingly, the budget shortfall was now taken care of, and the university was not required to put in any extra capital money over the initial $7.3 million (see table 1), which was very close to the savings from the USask contribution of $1 million a year toward the power bill of the old SAL facility, which was shut down from 1998 to 2005. USask did contribute a lot of important “in-kind” work from Matt Webster and Beryl Lepage, and a significant amount of money for UMA staff and the Mckercher Mckercher and Whitmore lawyers (Dick Batten and Doug Richardson, mentioned previously). For example, Matt Webster and Dick Batten certainly did a lot of good work to finalize the contracts with the province, the city, WED, and NRCan, among others (see table 1). All of these contracts held up very well. There have been no significant problems with any of them in the last eighteen years, as far as I know.

My Role as Interim Director, 1999–2001  93

c)  Getting an Agreement between the Owner, USask, and CLS Upon my arrival in Saskatoon in September 1999, Brian Hansen had been replaced by very helpful provincial officials such as Larry Spannier. CLS was to be owned by a single entity, USask; and by then, fortunately, there was no obvious tension between CLS and the province. But there were still many issues of disagreement between the owner, USask, and the fledgling CLS Inc. Dick Batten from Mckercher Mckercher and Whitmore had already drafted the articles of incorporation for the new non-profit corporation CLS Inc., and it was registered in the province of Saskatchewan on May 14, 1999. In the twenty-one-page document, many important issues were discussed, such as the Board of Directors and its many committees. In addition, Batten had already produced a draft operating agreement between USask and CLS, but this was far from agreement. One of my first meetings, on September 28, 1999, was with the university’s Business Committee to discuss this operating agreement. Remember, this was a new structure with very few examples to show us the way. Potentially, we still had another two-headed monster, with the CLS Board and the owner, USask, in constant conflict. It was surprising to me that my opinions on many issues concerning the management and control of CLS diverged from those of some of the USask administration. Several of these issues were so contentious that this agreement was not signed until three years later, in October 2002, after long negotiations involving mostly Dick Batten, Mark de Jong, CLS machine director and project leader, and Tony Whitworth, vice-president administration of USask, and his staff. Fortunately, the final agreement was reasonable, and the management and control issues were mostly solved. But, as was perhaps to be expected, there were some difficult, onerous, and abrasive meetings over this. From 2002 to 2017 (when I left the CLS Board), there were relatively few problems between CLS management, the board, and the university, with the university helping CLS in a number of ways. For example, CLS has continued to use the USask financial system, including for issuing all cheques. Some of the issues from 1999 to 2001 were trivial, while others were profound. The former included such questions as who was to sweep the floor and maintain the heating and air conditioning. Should the USask logo or name be included in all CLS communications? The more profound questions included the following: Who was responsible for negotiating all the contracts? Who was responsible for health and safety? The first part of this last question obviously arose during the above-mentioned negotiations with Ontario and Alberta. CLS managed to maintain control over these two contracts, and a number of other smaller contracts from ten other donors. The health and safety issue was a critical one. It was necessary to get CLS licensed with the Canadian Nuclear Safety Commission (CNSC) because of potential radiation problems. A lot of meetings (mainly involving Mark de Jong and various university officials) did not

94  The Canadian Light Source

solve the problem. At a meeting of the President’s Committee in 2000, Tony Whitworth declared that the university would be responsible. Mark de Jong and I were certainly taken aback, and said that CLS had to be responsible because it was operating the accelerators and had the expertise to do so. Moreover, the CNSC would insist on the operator (CLS) controlling health and safety issues. The senior CLS staff such as Mark de Jong and myself were not even on Whitworth’s organizational chart, and he had the university’s health and safety officer looking after the CLS health and safety portfolio from his USask office. President Peter MacKinnon, the committee chair, called for a break in the discussion for fifteen minutes, met with the university administrators outside the room, and came back and announced that CLS would be responsible for health and safety. What other president would chair such meetings in the first place, let alone make such difficult decisions involving his close colleagues so quickly? As noted above, the discussions with the university went on for another two years; and the USask-CLS agreement was reached after I left Saskatoon in October 2001. The final licence agreement between USask and CLS (twenty-five single-spaced, small-type pages) can be summarized as follows: USask has ­licensed its state-of-the-art third-generation synchrotron light facility to the new company (CLS Inc.), which is responsible for the operation and conduct of all activities related to the facility: its operation, performance, maintenance, hiring, and health and safety requirements, including the design, installation, operation, and servicing of all beamlines and related equipment. The fundraising role was to be shared between CLS and USask without any mention of lead roles. This was not a problem with Peter MacKinnon as president of USask. He spent a great deal of effective time in Ottawa helping CLS directors Bill Thomlinson and Josef Hormes obtain the first CLS operating grants in 2004 and 2009, respectively. However, recently the lack of definition of roles has again become a source of tension between the CFI, USask, and CLS, who apparently cannot agree on who has the lead fundraising role. d) Setting Up the Basic Organization within CLS, the Committee Structure, and Principles of Operation Mark de Jong and I had to begin organizing the employees into the usual groups in synchrotron facilities in other countries: the machine group to build the accelerators, the beamline group to build the beamlines, the industrial group to market the facility to industry, along with ancillary groups such as engineering, financial, outreach, health and safety, etc. Mark, who had worked in accelerator physics at Chalk River, was key to most of the initial hirings, and the number of employees grew, in a well-defined structure, from twenty-five in 1999 to sixty-five in 2002.

My Role as Interim Director, 1999–2001  95

But of equal or greater importance was the development of a committee structure to ensure good communication with the large number of employees, users, and funding groups, followed by development of the principles for financial contributions and operation. Several very appropriate and well-organized committees were established by Tony Whitworth in the summer of 1999, before I arrived in Saskatoon. Indeed, USask has to be congratulated on the effort expended by many dedicated administrators since 1993, who spent a very large amount of time on CLS committees, as well as representing CLS in Ottawa, exemplified by Peter MacKinnon’s efforts to obtain the operating funding in 2004 and 2009. The initial CLS Board was chaired by the NRC president, Arthur Carty, a chemistry colleague I had known and admired for many decades. Immediately, in September 1999, board members (all unpaid, with expenses covered) were appointed from the major funders and synchrotron users: from USask (Michael Corcoran and Tony Whitworth), WED (Doug Maley), the Saskatchewan government (Larry Spannier and Fraser Nicholson), University of Alberta (Ron Cavell), Western (Bill Bridger), along with the dedicated synchrotron scientist Adam Hitchcock from McMaster, and Dennis Skopik formerly of SAL. Walter Davidson from NRC, Dick Batten from Mckercher Mckercher and Whitmore, and Mark de Jong from CLS were often present as observers. There were four meetings per year (in March, June, September, and December) beginning in December 1999. These were organized extremely well in advance by Beryl ­Lepage and her staff, with very detailed minutes taken by a recording secretary. After my term as director from 1999 to 2001 (during which I was, of course, presenting CLS material to the board), I was very pleased and honoured to sit on the board from 2001 to 2005 and again from 2011 to 2017. During those last six years, the energetic and very involved chairs have been Walter Davidson and Nils Petersen. I believe that the board has functioned well over the last twenty years (with over seventy meetings), with many well-qualified academic, industry, and government members having four-year (sometimes extended to six-year) terms. Board information, often consisting of well over a hundred pages of information, was prepared extremely efficiently by Beryl Lepage and her staff. This was distributed before the meetings to the board members and several board subcommittees. Inevitably, many difficult issues arose, which were usually addressed in a reasonable and cordial way. There appeared to be rather little tension between CLS and USask – at least up to 2017, when fundraising and branding issues started to cause problems. There was initially an overarching President’s Committee (mentioned above) chaired by Peter MacKinnon, the new USask president, to approve very large purchases and oversee policies, among other duties. This committee included Michael Corcoran, Tony Whitworth, Arthur Carty, Walter Davidson, Larry Spannier, Dennis Skopik, Doug Maley, Ron Cavell, and Adam Hitchcock. Often Mark de Jong was present to get approval for large purchase items.

96  The Canadian Light Source

The Business Committee, whose duty it was to approve small purchases and monitor contract negotiations, was chaired by Vice-President Administration Tony Whitworth. Barry Hawkins and Martin Heikoop from the engineering firm UMA were already onsite to oversee the building construction and all contracts, together with Mark de Jong from CLS; along with Beryl Lepage, they worked very effectively together to ensure that the facility was built on time and on budget. The Industrial Committee, chaired by Vice-President Research Michael Corcoran, was created to try and identify industrial opportunities for CLS that would generate a significant percentage of the overall annual revenue. Then we also created oversight committees such as the Facility Advisory Committee (FAC) chaired by Alex McAuley, the Review Oversight Committee (ROC) chaired by my former student John Tse, the Machine Advisory Committee (MAC), the Users Advisory Committee (the UAC), and the Science Advisory Committee (SAC) (see references 2 and 3 for my articles on these various endeavours). All of these committees functioned relatively well. In response to the President’s Committee, Beryl Lepage and I spent a lot of time developing principles for contributions and beamline construction and operation, some of which are described below. e)  Planning for the Initial Phase 1 Beamlines It was essential to have a committee of scientists to choose which beamlines would be built; and then to select the principles and rules for the building, operation, scheduling, and control of the initial suite of beamlines, and then future beamlines. Emil Hallin was enormously helpful and important in this area. I chaired a Beamline Planning and Access Committee (BPAC) that met forty times over two years with detailed minutes. The BPAC’s main function was to approve the initial suite of beamlines. The committee was mainly composed of the key beamline developers who were experienced synchrotron users. Because each beamline had a different group of scientists heavily involved in its development (so-called beamline teams, or BTs), each with different ideas for the control and use of the beamline, each beamline had to be treated differently within a framework of principles. I have already mentioned the control issues between the province and CLS, and between USask and CLS. The BTs added yet another possible area of struggle for control, this time within CLS. One can only imagine the enormous difficulties that a separate beamline ownership entity such as Beam Inc. would have faced to get the beamlines selected, funded, and built. Beryl Lepage and I produced two documents that developed the important principles for beamline development, operation, and construction: “Canadian Light Source: Framework for Contributions” (discussed above), and “Beamline Development and Operation: Principles and Process.” These documents dealt

My Role as Interim Director, 1999–2001  97

with the vision and mission statements of CLS; the control and ownership of all equipment by CLS; the kind of user partnerships that would be allowed; the privileges (e.g., priority access) that the BTs would be provided with during the building of the beamline and after commissioning; the nature of the peer review for user proposals; the amount of industrial use allowed; and the industrial fees. There were many animated discussions over close to two years, at least five drafts of both documents were produced, and the documents were approved by many committees, including the BPAC, the UAC, the President’s Committee, and the CLS Board. I am not sure what happened to these documents when I left in October 2001; but they were important in organizing the facility in perpetuity. Although arduous, there was always good, open discussion among the dedicated Canadian SR users on the committee (Adam Hitchcock, Ron Cavell, De-Tong Jiang, John Tse, along with Dennis Johnson and Emil Hallin), and the overall results were really good, I believe. It turned out that we could only build seven of the proposed suite of nine beamlines because of financial problems with the overall project. As expected, the cut of the two beamlines resulted in a few irate users, including my excellent former graduate student John Tse, who was by now a Canada Research Chair at USask. In addition, the BTs, formed from some of the major beamline proponents, often had very different views on the amount of control and priority access that they felt they deserved for their considerable efforts. For example, one BT did not want any priority access for their group, while another requested 50 per cent priority access. CLS needed to give priority access to BTs because of their heavy involvement in beamline design and experimental programs; but CLS needed to control the amount of BT priority access so that general academic and industrial users would have significant access. The initial seven BTs provided no capital or operating funding for their beamlines.* Thus, 50 per cent priority access for a few scientists on one beamline was not reasonable from my point of view, even though this BT was the most involved technically and scientifically of any other BT. Also, the scientists hired to build the beamlines were hired by CLS in consultation with the BTs, and this sometimes resulted in conflicts when the BT did not want a particular CLS employee working with them on the beamline. A negotiated beamline memorandum of understanding was written for each beamline, outlining the percentage of overall beamtime allotted to the BT (priority access), and financial and technical responsibilities for the beamline operation. Most of these agreements were signed by the leader of the BTs and me very quickly; but, in the one case noted above, there was constant disagreement over the terms in many versions of the agreement. I would not sign the one agreement when I was director, and it lingered for months with the new directors, Mark de Jong * The BTs did raise money for nearly all later beamlines.

98  The Canadian Light Source

and Bill Thomlinson, before consensus was reached with the director of research, Tom Ellis, in 2002. This particular disagreement still festers today. Such conflicts reflect the often extremely difficult sociology of beamline operations at synchrotrons, with a large number of diverse scientists with very different viewpoints on beamline management and control. f)  Encouraging USask to Hire Key Synchrotron Users I thought that I had to help USask hire a number of faculty to use CLS. The facility, which should have over a thousand users after ten years of operation, simply could not be successful with only the existing two local synchrotron users, Louis Delbaere and the soon-to-be retired Wilson Quail. The university had promised six synchrotron-related faculty appointments in its applications, but of course the nature of these appointments was not specified. Were these appointments “on top,” with the money coming from the general university budget? Could they be appointed very soon? Or were they to be replacement appointments meted out over several years? What departments would get these appointments and how would they be funded? It was not easy to encourage the university to hire faculty that would use CLS. This problem was not unique to the synchrotron at USask. Other international universities with synchrotrons had big problems with this issue as well. For example, the University of Wisconsin, at its Aladdin facility, never did address this properly; and the University of Berkeley, at the ALS facility, also struggled for many years with it. And these were much larger, and much better funded, universities than USask. In November 1999, I arranged to take a few USask department chairs to meet with Vice-President Academic Michael Atkinson to discuss what the university planned to do. Peter MacKinnon had just started as university president, and I did not want to bother him initially. I pointed out to Atkinson that the USask-CFI contract showed that the university would hire six faculty that would be users of CLS. I also pointed out that those six positions were a very small part (less than 1 per cent) of the more than eight hundred total faculty, and that they should be “on top” appointments generated in the next four years before CLS became operational in 2004–5. These faculty positions would have to come either from the existing faculty allotment or be generated by new money that was not identified. Atkinson was taken aback, and we were told in no uncertain terms that no such appointments would be made in the short term; as Atkinson put it, “you will have to live in this university.” And indeed, no appointments were forthcoming in the short term. This meeting pointed out the difficulties at USask (and at most universities) with selective investments and budget prioritization, and Atkinson had just begun a very necessary prioritization procedure at USask.12 As MacKinnon points out in his book, such

My Role as Interim Director, 1999–2001  99

procedures were not met enthusiastically by many areas of the university – a very significant understatement.12 Small rearrangements of the university budget in the previous year had obviously led to considerable collective angst, which Atkinson reflected in his negative reaction to my request. By April 2000, the university established a committee to examine if and how new faculty could be appointed that would be heavy users of CLS. Surprisingly, perhaps, I was not asked to be on the committee, and there was no one on the committee who had come close to a synchrotron or the use of SR. I was eventually asked to go to one of the meetings. I knew the departments that should receive SR appointments, and was quite clear about this at the meeting. However, a senior member of the committee became very annoyed at me, and directly accused me of not doing my job at USask to help them solve their “problem.” This outburst was especially frustrating because the committee member had not invited me to give a talk to his department, which became a major user of CLS. Fortunately, the committee member apologized after the meeting. But I decided that night that I did not need that kind of unreasonable conflict, and would not consider staying on as director after my two years as interim director. There were just too many good reasons to return to London: my son and daughter, my talented and friendly students and colleagues at Western, and my many friends and activities. To try to solve the problem of very few users in Saskatchewan, there were serious efforts made to establish a synchrotron centre or institute, the aim of which would be to aid in hiring talented faculty in many areas and to coordinate and enhance collaborative research across many academic and industrial areas for the benefit mainly of Saskatchewan users. As far back as March 1998, USask and the University of Regina had signed a memorandum of understanding to establish the Saskatchewan Centre for Synchrotron Research. Plans for an equivalent Saskatchewan Synchrotron Institute (SSI) began in the summer of 2000, and founding member organizations were the province of Saskatchewan (through the Saskatchewan Economic and Co-operative Development, or SECD), the Government of Canada (through WED/NRC), USask, and the University of Regina. A very good committee was formed, which included Mark de Jong, Rob Slinger, and myself from CLS, Doug Maley and Ed Wiens from WED, Wayne Craig and Bill Smith from NRC, Katherine Bergman from the University of Regina, and Larry Spannier, Anne Broda, and Dale Botting from SECD. At least seven business plans were drafted between October 2000 and July 2001. In July 2001, a mission statement was agreed upon: “to develop, coordinate and deliver strategies and programs designed to maximize the economic and scientific benefits to Saskatchewan associated with the construction, operation, growth and development and research opportunities of the CLS Synchrotron project.” In the seventh business plan, the budget for the institute was just $1.4 million for three years, but there were plans to leverage further funds to help support new faculty appointments.

100  The Canadian Light Source

I tried very hard to encourage the formation of SSI, for example, by arranging a visit to Stanford University and the Stanford Synchrotron Institute in the summer of 2000. One of my international colleagues from Stanford, Gordon Brown, a member of the CLS FAC, very generously offered to show President M ­ acKinnon and all his vice-presidents what Stanford had done with a synchrotron institute to attract outstanding faculty. With MacKinnon and the vice-presidents (except for Vice-President Academic Michael Atkinson), we went to Stanford and met with the school’s president and the Stanford synchrotron group. President MacKinnon became very keen to put university money into establishing an institute. Such an institute or centre was the best way to attract excellent faculty CLS users to USask, and it would ensure that USask would benefit more from other national and international users. On flying back from San Francisco, MacKinnon indicated that he would consider putting a few million dollars into such an institute. But his colleagues obviously talked him out of such an important initiative, and I heard nothing more about it from anyone at the university. It soon became obvious that USask would not finance the SSI, and the institute concept became a reality in 2002–4 in a very limited form (page 92). I did not approach MacKinnon again about it, but the lack of USask enthusiasm for the SSI confirmed that I would not consider staying as director after September 2001. Gordon Brown resigned from the FAC after realizing that USask was not going to establish a synchrotron institute. By the end of 2000, two good synchrotron appointments had been made by departments at the USask, but the real problem of too few users had still not been addressed. Coincidentally, the federal Canada Research Chair (CRC) program was announced in December 2000. It enabled all Canadian universities to appoint junior and senior chairs ($100,000 and $200,000 annually for junior and senior chairs, respectively) in numbers proportionate to their tri-council (NSERC, SSHRC, and MRC) research income. For example, Toronto received over 200 chairs, while USask received 33 initially and Western about 65. Shortly after, USask received up to 40 chairs, which settled back to 35. This made it much easier to assign some of these CRC appointments as CLS users. I believe that Peter MacKinnon ensured that ten of these would be for CLS users – a very good number from my perspective. I am sure this was very difficult for him, given university politics. MacKinnon’s decision was undoubtedly linked to the identification of “synchrotron sciences” as one of the six areas of strength at the USask that would receive priority support and enhancement. Indeed, ten research faculty using CLS in many different disciplines were appointed to CRCs by 2005, and some of these have generated outstanding research at CLS. For example, John Tse (physics), Louis Delbaere (biochemistry) and Dean Chapman (anatomy and cell biology) had CFI-related chairs initially; and as of 2019, Miroslaw Cygler (biochemistry), Graham George and Ingrid Pickering (environmental sciences), Ajay Dalai (chemical engineering), Alex Moewes (physics), and David Cooper (anatomy and cell biology) hold

My Role as Interim Director, 1999–2001  101

CRCs. In addition, many other newly appointed CLS users across the campus (e.g., Andrew Grosvenor in chemistry and Derek Peak in soil science) have generated excellent research at CLS. The problem from my point of view was that USask never used close to its allotment of CRCs. For example, USask never appointed more than 28 chairs, and as of September 2019, just 19 chairs are listed on the USask website. USask lost over 10 chairs that were initially allotted to the university, and it appears the school is about to give up even more. It does not seem reasonable to turn down between $1 million and $2 million a year for close to fifteen years. From my detailed analysis of CRCs at other Canadian universities, only one other institution gave up so many chairs. These additional 10 chairs at USask could have been used to enhance other areas of the university, not just CLS. In addition, universities then received “indirect cost” income, about 20 per cent of their tri-council research income, which amounted to between $2 million and $3 million just from some of the federal funding (NSERC and MRC) used to match the CFI funding to CLS. Very generously, USask returned nearly all that income to the CLS operating funds; but some of that money could have gone to enhance other parts of USask as well. I left Saskatoon and CLS in early October 2001, but remained as acting research director for four more years, working half-time for CLS while based at Western. The most important part of this job was to coordinate the grant applications submitted in May 2003 to the CFI for seven phase 2 beamlines and an experimental chamber.34 Three of these 8 applications came from USask as the leader, 3 from Western, and 2 from UBC (at a total cost of $55 million), but they were coordinated by CLS on behalf of USask. These beamlines opened up many new and exciting research opportunities to support existing pre-eminent research in agriculture, health, environment, and advanced materials. All of these were large applications, with letters of support raised from twenty-seven Canadian universities and another ten organizations. This exercise was again complicated greatly by the necessity of raising 60 per cent of the budget from organizations other than the CFI, which again provided 40 per cent of the money. Many supporting letters from the lead universities and provinces indicated their scientific and financial support for 40 per cent of the project costs. Six of the beamline proposals were successful, and were built successfully by 2009. Dennis Johnson continued working for CLS in advisory capacities until 2004, was director (as mentioned on page 92) of a reconstituted SSI from 2002 to 2004, and was largely responsible for attracting a few CRCs. But, unfortunately, this institute did not continue after operation of CLS began. Overall, the two years that I spent in Saskatoon, though extremely challenging, were very rewarding. It was amazing to see the construction of the very large and complex building, with terrific leadership from Mark de Jong of CLS and Barry Hawkins and Martin Heikoop from UMA. President Peter MacKinnon, Dennis Johnson, Doug Richardson, and board chair Arthur Carty were

102  The Canadian Light Source

very supportive throughout those two years. Many employees did yeomen jobs, some described above. For example, Beryl Lepage was seconded to finance at CLS from USask, and was a great addition. During my time in Saskatoon, she was responsible for overseeing the Ontario and Alberta contracts, the operation and framework documents, and the many board meetings, to name a few activities. She and her staff controlled the very complex capital and operating budgets. Because of the multiple funders of CLS, this was a difficult task. This problem was especially acute for the operating budgets, mainly because the larger funders (e.g., NSERC and CFI) have a variety of different spending regulations attached to their grants. In addition, after 2005, the financial group had to cope with significant cost overruns on the phase 2 and phase 3 beamline projects. It has long been recognized that this multiple funding model for such large diverse national facilities is not satisfactory; but so far there is no sign that either the CFI or the federal government will change to a single funder for the operating grant to CLS. It was very pleasing to hire a few of my old students, such as Brian Yates and Jeff Cutler, who became critical for beamline construction and the industrial business, respectively (see figures 7 and 9). Several other former students and post-docs at the Madison CSRF facility – De-Tong Jiang, Kim Tan, Yongfeng Hu, Emil Hallin, Ian Coulthard, Tim May, Jigang Zhou, and Narayan Appathurai – were also hired by 2005 (see figure 9). Other former Western graduate students of mine, John Tse, Grant Henderson (University of Toronto), and of course T.K. Sham, became very productive CLS users as well. There were many talented and dedicated people in the administration at the facility: beginning with Lavina Carter and Angela Shenher, who both did so much excellent administrative work at SAL with Dennis Skopik and then with CLS, and Beryl Lepage, who has been referred to previously. Joan and I became active in the Saskatoon community. For example, Joan took theology courses toward her master of divinity at St. Andrew’s College at USask in the first year, and then took a very rewarding ministerial job at Mayfair United Church, with the senior minister Jack Carr, in the second year. Curling, golfing, and baseball were important to meet and get to know many of the staff. Dennis Johnson, Jeff Cutler, and Rob Pywell from the physics department curled with me in a competitive league for the two winters. It was wonderful to be able to visit several times with my ninety-five year-old favourite aunt. She lived to 102. The road between Regina and Saskatoon was well travelled. Probably my biggest thrill was the official opening of CLS in September 2001, about a month before I left the CLS director’s job. Prime Minister Jean Chrétien and representatives from the funders and researchers were present (see figure 11). The huge building was awesome, and the two large circular accelerators were beginning to take shape on the floor.

My Role as Interim Director, 1999–2001  103 Figure 11.  The official opening ceremony of CLS in September 2001. Left to right: Peter MacKinnon (president, USask), Mike Bancroft (director, CLS), and Jean Chrétien (prime minister of Canada).

The accelerators and beamlines took four more years of development before the first experiments were conducted at the end of 2005. On October 22, 2004, CLS officially opened with electrons circulating in the synchrotron, but the first experiments on the phase 1 beamlines were conducted toward the end of 2005. The October 2004 opening ceremony was attended by many university presidents, government officials, and leading scientists, including the federal minister of finance, Ralph Goodale, the premier of Saskatchewan, Lorne Calvert, USask president Peter MacKinnon, the CLS’s director (from 2002 to 2008), Bill Thomlinson, and the chair of CLS Board, Arthur Carty. Peter Mansbridge hosted The National, the CBC’s nightly newscast, from the top of the storage ring on October 21, 2004. The seven phase 1 beamlines were completed by 2006 on budget, and this process was led by the CLS beamline developers Emil Hallin, Brian Yates, Tim May, De-Tong Jiang, Ning Chen, Yongfeng Hu, Ian Coulthard, Konstantine Kaznatcheev, and Pawel Grochulski. These individuals did an outstanding job, and they benefited from all the technical and engineering help provided by so many at CLS and the beamline teams. The five phase 2 beamlines were also

104  The Canadian Light Source

implemented expeditiously by 2009. However, the seven phase 3 beamlines, proved to be a greater challenge, and have been seriously delayed. (Technical and scientific details of all these beamlines are given in a very recent article; please see reference 34.) All of these large projects (costing from $5 million to $25 million each) have entailed significant personnel, financial, and technical challenges, which have led to very long delays and much frustration. But with substantial reorganization of the staff and strong leadership from the present director, Rob Lamb, these beamlines are now operational and have been nearly commissioned. More recently, funding for three beamline upgrades has been obtained (so-called phase 4). The details of this period, lasting from roughly 2005 to the present, should be written in a separate book in the near future.

7  The CFI: Goals, Impact, and Paul Martin

a)  Total CFI Funding to CLS As indicated above, the Canada Foundation for Innovation was the largest single funder of CLS. As detailed in table 1, the CFI contributed the crucial $56.4 million in 1998 to initiate the facility (the total cost was 140.9 million). Since then, the CFI has made very significant investments (more than $40 million) to CLS to fund the construction of fourteen new beamlines – so-called phase 2 and phase 3 beamlines34 – and the purchase of a substantial amount of equipment. The CFI has also contributed a large fraction of the operating expenses in the last few years. CLS, as a national facility, had a great advantage in the many subsequent CFI competitions, which were often focused on only a few national facilities (see table 2). As of March 2017, the total CFI funding to CLS (capital plus operating) since 1999 was $215 million ($105 million in capital and $110 million in operating); while the total funding to CLS is in excess of $650 million since 1999 (more than $300 million in capital and $350 million in operating). It is important to note that $175 million (as shown in table 2) of the $215 million worth of contributions made by the CFI accrued to USask. Most of the remaining $40 million accrued to Western, UBC, and Guelph, who applied for several of the phase 2 and phase 3 beamlines and equipment. The co-operation between all the universities was good in this very complex environment. The CFI has contributed close to 30 per cent of the total funding to CLS and most of the rest of the capital money came from the major applicants’ home provinces. As shown below, these amounts are a major reason for the very large increase in research funding to USask, which in turn sparked a corresponding increase in the university’s reputation. USask has also been very successful with major grants to VIDO (now VIDO-InterVac), a level-three microbiology research facility (see table 2), and two (out of thirty national) new Canada Research Excellence Chairs in Food Security and Water Resources. The CRC program has also contributed close to $20 million to USask for the ten chairs that use CLS.

106  The Canadian Light Source Table 2.  Large CFI grants to CLS and VIDO (1999–2016) CLS

VIDO

Date

Amount $M

Date

Amount $M

March 1999 March 2004 November 2006 March 2012 September 2016  Total

56.4 6.8 8.3 55.2 49.0 175.7

July 2000 March 2004 September 2016

5.2 32.5 19.3

 Total

57.0

Source: From the CFI website

b)  The History and Importance of the CFI Since 1999, CFI and CRC funding has made a huge impact on research capacity all across Canada, as illustrated in table 3. Close to 150 universities, colleges, and hospitals all across the country have received over 10,000 grants totalling over $7 billion. During this period research funding has at least tripled at a number of Canadian universities (see the ratio column in table 3), and the overall research funding has close to tripled at other institutions as well. USask has more than quadrupled its funding since 1999, mainly because of CLS and VIDO (see table 2), and this will be examined more closely in chapter 8. This enormous increase in Canadian research funding is due mainly to the formation of the CFI, which not only provided a significant proportion of the increase, but forced a change in the research “climate” within all provinces, at most universities, and among many Canadian academic researchers. The CLS facility, along with many other new facilities and their associated academics across Canada, owe a large debt of gratitude to the CFI and the Liberal Party, and in particular to Paul Martin and Jean Chrétien. Because of the central importance of the CFI in funding CLS, it is critical to emphasize its history and mandate.* The finance ministry, headed by Martin and ably supported by many dedicated bureaucrats such as Scott Clark, deputy minister of finance, set several priorities in early 1998 after they had slain the deficit and surplus money began to be available. They felt that, in an extremely wealthy society such as Canada, the highest priority was to “level the economic playing field” for all Canadians. Income disparities were increasing, and that was not compatible * Some of this section draws heavily on my telephone interview with Paul Martin in early 2010.

The CFI: Goals, Impact, and Paul Martin  107 Table 3.  The total research support in 1999 and 2016 ($M) for universities with over $30M funding in 1999, with the ratio of the two figures University

1999 ($M)

2016 ($M)

Ratio: 2016/1999

Toronto Montreal McGill Alberta British Columbia Calgary Laval Western Ottawa Guelph McMaster Queen’s Manitoba Waterloo Saskatchewan Dalhousie Sherbrooke Quebec (Mont.)

306.5 (1) 206.2 (2) 198.9 (3) 174.3 (4) 139.1 (5) 107.8 (6) 96.8 (7) 94.9 (8) 89.3 (9) 88.8 (10) 85.9 (11) 68.9 (12) 67.7 (13) 57.3 (14) 51.1 (15) 45.0 (16) 33.7 (17) 30.1 (18)

1,008.3 (1) 522.9 (4) 547.5 (2) 433.7 (5) 532.1 (3) 360.5 (7) 376.9 (6) 234.7 (10) 326.0 (9) 148.9 (15) 354.6 (8) 151.8 (14) 191.3 (12) 166.4 (13) 215.9 (11) 136.0 (16) 129.2 (17) 63.9 (18)

3.29 (8) 2.54 (13) 2.75 (12) 2.49 (14) 3.82 (5) 3.34 (7) 3.89 (3) 2.47 (15) 3.65 (6) 1.68 (18) 4.13 (2) 2.20 (16) 2.83 (11) 2.90 (10) 4.23 (1) 3.02 (9) 3.83 (4) 2.12 (17)

Note: The order of universities is given by the 1999 funding. The ranks are given in brackets in all three columns. Source: RE$EARCH infosource, http://researchinfosource.com/pdf/top50.pdf (November 16, 2017)

with their philosophy. They also felt it was critical for Canada to “advance the state of human knowledge” in all areas, and that basic science had to be well supported as a base for economic growth in Canada. Moreover, they felt strongly that the government had a responsibility to act on these two priorities. It is worth looking now at the finance department’s efforts in the second area, that of scientific research. As an aside, it strikes me that it is extremely unusual for any federal finance department to be leading the charge on such a research initiative. It would be interesting to investigate any similar initiatives in other countries. In the mid-1990s, there was strong lobbying of the federal government by many scientific groups – for example, NRC, MRC, the Canadian Society for Chemistry (CSC), and the Canadian Association of Physicists (CAP) – to increase federal funding for science. Indeed, I was part of that lobbying effort as the president of the CSC in 1996–7. Because of the deficits in the 1980s and ’90s, research funding had been greatly restricted during those years. Canada was becoming less competitive in many research areas because of difficulties in attracting outstanding researchers and the lack of state-of-the-art equipment.

108  The Canadian Light Source

But how would the federal government support a large increase in research funding? For starters, it could have increased funding to the three Canadian research councils: NSERC, MRC (now the CIHR), and the Social Sciences and Humanities Research Council (SSHRC). However, the government wanted to set up a fund independent of government, effectively an endowment fund that would not be affected by annual budget fluctuations or be subject to political control. It was apparent that this concept was scrutinized strongly by the auditor general; but within a few months of eliminating the federal deficit, the CFI was formed and was ready for applications for approximately $200 million in the phase 1 competition. The main goals of the CFI initially were to strengthen the research and development (so-called R&D) capabilities of universities and institutions (such as hospitals), and encourage collaboration with industries. Together with the CRC program, the CFI sought to stop the “brain drain” of Canada’s most talented academics, and to attract outstanding international talent by providing the very best research equipment and the best collaborative opportunities. The CFI’s regulations insisted that the foundation would fund only 40 per cent of the total capital (and later operating) expenditures, with the other 60 per cent coming from the provinces (40 per cent) and industry (20 per cent). These regulations were intended to leverage a lot more money from the provinces, cities, industries, and universities; and indeed, the 60 per cent was usually raised. Every province in Canada established an organization to ensure that their successful applicants could obtain the usual 40 per cent provincial matching money. The final 20 per cent usually came from industrial and other contributions. For example, Ontario established OIT, which was directed by David Bogart, who I talked to initially in Toronto in September 1999 (see chapter 6). All major universities immediately created offices under their vice-president research to prioritize applications within their institutions, to help write the grants of choice, and sometimes to help to identify the remaining 20 per cent matching money. These procedures forced nearly all universities to expand greatly their vice-president research offices, to prioritize their areas of support, and to selectively enhance chosen areas of excellence with CRCs. The CFI also greatly encouraged large, multidisciplinary collaborations between many academics at many different institutions – a difficult and controversial exercise at any university, let alone across multiple institutions. Of course, well before the CFI was formed, CLS was a marvellous example of just such a multidisciplinary academic collaboration to create a truly national facility. Indeed, it is very likely that CLS was at least partly responsible for the formation of the CFI (as seen in chapter 5), and probably was important for the conceptual and philosophical basis of the entire CFI program. USask quite naturally made synchrotron sciences one of their six areas of excellence for future enhancement.

The CFI: Goals, Impact, and Paul Martin  109

All of the above led to a profound change in attitude toward research at many universities in Canada. The CRC program also produced a huge fund for the enormous benefit of the entire Canadian university community, including the non-scientific disciplines. To address the overall health of the universities, most institutions insisted that a significant portion of their CRCs should be awarded in non-science areas – such as social sciences and the humanities – that did not produce a significant part of the tri-council research income. (As noted on page 100, this income generated the total number of CRCs for each university.) For example, as of January 2018, USask has three CRCs, in public policy, regional innovation, and Indigenous rights. The effect of the CFI program is readily apparent in table 3, which shows the large increase in research funding at nearly all universities during the 1999– 2016 period. Many universities more than tripled their research income during this time, with USask more than quadrupling its research income. Since 1999, there is no doubt that Canada has much more talented research personnel, and many more excellent up-to-date scientific facilities, such as CLS. The CFI method of funding (with only 40 per cent coming from the CFI) has been a very effective way of encouraging the provinces to increase their research funding. However, identifying the multiple sources of capital and operating funding requires a tremendous amount of professional time and effort. This approach is especially ineffective for raising operating funds, particularly for a facility as large as CLS. At least seven organizations contribute to CLS’s operating funds: CFI, NSERC, NRC, CIHR, WED, USask, and the province of Saskatchewan. These organizations all have different spending requirements and regulations. This is not a reasonable way to fund the operation of a large national facility. Hopefully, the great majority of the operating funds for CLS will come from one national government agency in the near future. There has been general disappointment in the increase in industrial spin-offs and economic impact in Canada resulting from the huge CFI investments, partly because of the difficulty in raising venture capital and other monies for new industries in Canada compared to the United States. For example, the formation of the important spin-off company Canadian Isotope Innovations (CII) Inc. in Saskatoon has been handicapped by such fundraising problems. Hopefully, CII will become a very major supplier of isotopes for medical imaging in Canada and abroad (see chapter 8). Because of the less-than-expected economic impact of the CFI funds in the last eighteen years, the ISED is now proposing to spend $950 million in up to five business-led innovation superclusters. Dr. Runte, president and CEO of the CFI, stated that “this initiative is a new opportunity to strengthen Canada’s most promising clusters, and to build superclusters at scale. This is a cooperative effort where the private sector, academic institutions, not for profits,

110  The Canadian Light Source

SMEs and our government come together to build the economy of the future and create resilient well-paying jobs.”* The superclusters from industry-led consortia are in sectors ranging from ocean technologies, to artificial intelligence, to advanced manufacturing and agri-foods. Hopefully, these will yield a much better return for the overall economy in the future. One of these superclusters announced in February 2018, Protein Industries Canada, is centred in Saskatoon, and CLS has a role within it.

* “Excellence in research and innovation,” supplement in The Globe and Mail, November 21, 2017.

8 The Positive Impact on USask and Canadian Science

a)  Overview of CLS operation Before describing the impact of CLS on USask and Canadian science more generally, it is important to briefly outline the current organization of the facility along with the number of users and some statistics. CLS is now a very large organization with an operating budget of over $30 million a year, a few dedicated senior managers (e.g., the director and associate directors), and about 250 employees. These employees are mostly in the following seven groups. First, physicists, mechanical engineers, and computer experts are required to design, build, and maintain the electron accelerators and beamlines. Second, physicists and chemists operate the beamlines, generate research, and optimize beamline use for over a thousand users per year. Third, a financial group produces and controls budgets with the great complexities inherent in this multi-funding environment. Fourth, a users group coordinates the scheduling, reporting, and evaluating of applications from the thousand-plus users per year on over twenty beamlines, including a beamline at the APS in Chicago.* This beamline was initiated and designed by Daryl Crozier at Simon Fraser University. Fifth, an industrial group helps industries access the facility and generate applied scientific results with an important revenue stream. Sixth, an outreach group educates outside lay people, teachers, and students with tours, lectures, and simple experiments. And finally, a health and safety group oversees the radiation safety of the facility and reports to the Canadian Nuclear Safety Commission. The facility can accommodate many hundreds of users a year because all beamlines operate independently twenty-four hours a day, seven days a week * The CLS-APS partnership arose after a 20-year collaboration (initially to form a collaborative access team (CAT)) between Crozier and Canadian and US colleagues from the University of Washington and the Pacific Northwest National Laboratory. As a result, many Canadian scientists have greatly benefited, including gaining access to many other APS beamlines.

112  The Canadian Light Source

for the majority of the year. The synchrotron has provided over 5,000 hours of beamtime nearly every year to all the working beamlines, with the approximately 3,500 hours remaining scheduled for maintenance and shutdowns. A large number of scientists from across Canada and around the world visit CLS regularly for a few days’ research, usually once or twice a year. In the first ten years of operation, from 2005 to 2015, there were 2,584 users (amounting to over 10,000 user visits), 53 per cent of whom were graduate students. In recent years there have been over 1,000 users per year (table 4). The majority are from Saskatchewan, but there are many users from other parts of Canada (ten provinces and two territories) and around the world (twenty-eight countries). The facility has generated over 200 peer-reviewed publications per year for many years; but in the last few years, over 500 papers (more than 300 of them peer-reviewed) have been published. Users come from nearly every discipline in science, engineering, and medicine, along with archaeology and anthropology, as indicated by the twenty-seven departments at USask that have CLS users (see table 5). Many of the resulting papers have appeared in the most prestigious academic journals, such as Science, Nature, and Physical Review Letters, and many have been widely cited. (Examples of a few such papers will be given shortly.) It is also important to emphasize that most users will only travel long distances to such a facility if it produces unique results competitive with other synchrotron facilities in the United States and abroad, and if there is excellent technical assistance available. CLS represents a marvellous Canadian technological achievement based on over a century of modern physics, constructed on time and on budget with the co-operation of many national and international scientists working at CLS and abroad. To give some idea of the complexities of this facility, the developed area of CLS is 12,700 m2 (comparable to seven ice-hockey rinks). There are 900 billion electrons accelerating in the storage ring, with the electrons going so fast that they would reach the moon in one second. The cross-sectional area of the electron beam is less than the area of a human hair. CLS began operation in 2005 with seven initial beamlines optimized for high intensity and resolution in the far infrared, infrared, soft x-ray, and hard x-ray regions of the electromagnetic spectrum; and CLS’s capabilities have expanded, with eighteen more beamlines in the soft and hard x-ray regions since 2009. The main storage ring was designed in-house. In addition to being unique, it is arguably the most cost-effective SR source in the world. Other comparable facilities constructed in England, France, and Australia at a similar time were about twice as expensive. Moreover, the accelerators and initial seven beamlines were constructed on budget, without any requests for other funding in the five-year period covered by the initial $140 million budget (see table 1). The great success in obtaining funding for the many phase 3 beamlines caused serious delays in the construction of those beamlines, mainly because the technical staff was overcommitted. Also, there were great difficulties in obtaining the 60 per cent

The Positive Impact on USask and Canadian Science  113 Table 4.  Users per year by geographic distribution Geographic Distribution Alberta British Columbia Manitoba New Brunswick Newfoundland Nova Scotia Northwest Territories Ontario Prince Edward Island Quebec Saskatchewan Yukon   Total Canada Total International   Total Users

2012

2013

2014

2015

39 52 18 7 3 7

55 60 14 4 3 5

55 60 7 7

  112   23 329   590 165 755

  123 40 403 8 715 168 883

65 75 10 5 4 136 26 380 701 195 896

2016 63 52 13 7

 

  2

11

  146   34 359   670 197 867

197   47 471   861 233 1094

Source: CLS User’s Office, 2017. Table 5.  USask users by college, department, etc. Colleges, Schools

Department

Centres, Institutes, etc.

1. Agriculture, Bioresources

Animal and Poultry Science

Food Innovation Institute

Plant Sciences Soil Science Food and Bioproduct Sciences Archaeology and Anthropology Biology Chemistry

Crop Development Centre

2. Arts and Science

Geological Sciences History Physics and Engineering Physics 3. Dentistry 4. Education 5. Engineering

Biomedical Chemical and Biological Civil, Geological, and Environmental

Applied Microbiol.& Food Science Museum of Antiquities Sask. Structural Sciences Centre Plasma Physics Lab

Sask. Centre for Excellence in Transportation and Infrastructure (Continued)

114  The Canadian Light Source Colleges, Schools

6. Medicine

Department

Centres, Institutes, etc.

Electrical and Computer Environmental

Nanotechnology Environmental Sciences and Biotechnology

Mechanical Physics and Engineering Physics Anatomy and Cell Biology Biochemistry Medical Imaging

Saskatchewan Cancer Agency Division of Hematology and Oncology

Microbiology and Immunology Medicine (Neurology) Ophthalmology Pathology Pediatrics Physiology 7. Pharmacy and Nutrition 8. Western College of Veterinary Medicine

Veterinary Pathology Veterinary Microbiology Veterinary Biomedical Sciences Small Animal Clinical Sciences

School of Environment and Sustainability Others associated with USask

Prairie Swine Centre TR Labs (now TR Tech) Toxicology Centre Global Institute for Water Security Vaccine and Infectious Disease Organization (VIDO) International Vaccine Centre (INTERVAC)

matching funding. This caused major headaches for the CLS finance group, the director, and the CLS Board. But, the CFI has never been asked for extra funds for those projects over and above the original beamline allotments. Most importantly, the synchrotron source met specifications, most beamlines met specifications, and many of the beamlines are very competitive with comparable beamlines at other international synchrotron facilities. The

The Positive Impact on USask and Canadian Science  115

excellent performance of most of the beamlines is a fantastic tribute to the Canadian academics and CLS staff that designed and constructed the building, ­accelerators, and beamlines (see appendix 1); and also the CLS financial group headed by Beryl Lepage, which managed to control the complicated funding and budgets scenarios. Indeed, Canada has a proven track record of producing world-class, large-scale science facilities requiring extensive collaboration not only between domestic institutions and individuals, but also between international ­colleagues and agencies. Facilities such as TRIUMF in Vancouver and the ­Sudbury ­Neutrino Observatory in Sudbury, Ontario (whose director, Art ­McDonald, won the 2015 Nobel Prize in Physics for characterization of ­neutrinos) show that Canada can produce world-class facilities time and again. b)  Impact on USask CLS had only 2 users from USask in 1999, but that number increased dramatically to more than 470 in 2016 (see table 4); these researchers came from 8 colleges (or faculties), 1 school, 29 departments, 11 centres/institutes, and 7 large laboratories on or near the USask campus (see table 5). USask academics, and indeed the larger research community, have responded to the enormous opportunities provided by CLS; and the whole Saskatoon community (the university administrators and researchers, CLS management and employees, and the CLS Board) should feel very proud of their accomplishments. Looking back, the decision to have USask as the owner of CLS was probably better than having NRC assume this role. I really doubt that USask faculty and students would be as heavily involved with CLS if the facility was an NRC laboratory, and there would be many fewer USask student and faculty users. Also, university ownership minimizes operational disruptions at CLS that may result from changes in national science policy brought about by successive governments. The success of CLS has indeed made a large impact on the research profile of USask and its overall reputation in Canada. Even during the construction phase, President MacKinnon was clear, writing in the 2000–1 USask annual report that “CLS symbolizes the future of this university. With the completion of this world class facility, Saskatoon will attract leading students, researchers and industrial partners from across the nation and around the world.” The data presented below show that CLS has indeed fulfilled President Mackinnon’s prophecy. In 2001, CLS still had to deliver on its promises. But immediately, the facility’s funding, as shown in table 1, dramatically enhanced the USask research revenue. As we see in tables 6 and 7, for example, CLS provided the great majority of the dramatic increase in research revenue, from $51.1 million in 1999 to $71.1 million in 2000, and on to $100.5 million in 2001. Understandably, the university was eager to advertise in its 1999–2000 annual report that its “research revenues soared to $71.1M in 1999–2000. This is a 39% increase over the previous year’s revenues and the largest percentage rise in 25 years.” What was not mentioned was that

116  The Canadian Light Source Table 6.  Total USask research revenues by college/unit, 1999–2001 (M$) College/Unit

1999–2000 (partly)

2000–2001

Canadian Light Source Agriculture Interdisciplinary Medicine Arts and Science Other support Engineering Veterinary Medicine Graduate Studies Pharmacy Education Kinesiology Dentistry  Total

16.2 (23%) 15.2 (21%)

36.7 (36.6%) 15.5 (15.4%) 11.1 (11.0%) 10.7 (10.6%) 9.8 (9.8%) 5.5 (5.4%) 5.0 (5.0%) 3.5 (3.5%) 1.5 (1.5%) 0.9 (0.9%) 0.1 (0.1%) 0.1 (0.1%) 0.1 (0.1%) 100.5 (100%)

11.0 (15%) 8.1 (11%)

71.1 (100%)

Source: From the 1999–2000 and 2000–2001 USask Annual Reports

$16.2 million of the $20 million increase (more than 80 per cent of the total) from 1999 to 2000 came from CLS funding (see tables 6 and 7). Over the two years covered in table 6, CLS contributed $16.2 million plus $36.7 million (as shown in the top lines of table 6), for a total of $52.9 million of the total increase of approximately $70 million – or more than 75 per cent of the increase. Most importantly, this CLS funding elevated USask from fifteenth place to twelfth in terms of research funding in Canada (see table 7), and it provided a large incentive to prioritize and improve the research performance generally at the university. Since 2001, CLS equipment and operating monies have on average contributed more than $30 million a year to USask (see tables 3 and 7). Unlike the figures from the 1999–2001 period, these numbers are not available directly from the USask annual reports, but several USask annual reports state that research income from “other units” is greater than $30 million, and most of the funding to CLS (a total to date of $650 million minus $140 million – the initial funding shown in table 1) averaged over the years 2004 to 2016 is in excess of $35 million per year. It is important to comment that these annual reports show that CLS funding is the largest single contribution to the USask research budget. Indeed, in 2011 and 2012, the CLS contribution was larger than the total tri-agency (NSERC, CIHR, and SSHRC) revenue of approximately $30 million. So much for the initial statements from some faculty that CLS funding would have a very negative impact on many parts of USask! Instead, it has enhanced the school’s reputation greatly, such that it is now counted among the top fifteen universities in Canada. USask has used its increased funding to enhance its reputation in other, perhaps more important ways as well. For ­example,

The Positive Impact on USask and Canadian Science  117 Table 7.  A comparison of the total sponsored research funding (in $M) for USask and Western from 1999 to 2016, along with their rank amongst Canadian universities Year

USask $M

1999 2000 2001 2002 2003 2004 2005 2006 2007 2008 2009 2010 2011 2012 2013 2014 2015 2016

51.1 71.9 101.6 121.3 116.8 107.6 110.4 106.8 150.5 203.5 169.5 184.8 203.2 166.7 158.0 195.3 169.0 215.9

Rank 15 15 12 12 13 15 15 16 13 11 13 12 11 13 14 11 13 11

Western $M 94.9 109.2 131.8 149.3 145.8 191.2 179.9 225.9 238.0 222.3 241.7 221.2 218.7 241.1 254.7 237.9 229.8 234.7

Rank 8 9 10 11 11 9 10 10 9 10 9 10 10 10 10 10 10 10

Source: RE$EARCH infosource, http://researchinfosource.com/pdf/top50.pdf

USask was invited in 2011 to join what is known as the U15, the fifteen most research-intensive universities in Canada. Originally established in 1991 as the Group of Ten (Toronto, Montreal, McGill, British Columbia, Alberta, McMaster, Western, Laval, Queen’s, and Waterloo), the group was expanded in 2006 to include Dalhousie, Calgary, and Ottawa. President MacKinnon sought USask’s membership as an indication that the university was in the same league as the rest, and, along with the University of Manitoba, it joined the group in 2015. MacKinnon points out that with a stronger medical school, USask has every prospect of moving into the top echelon of U15 universities.12 As discussed below, several other areas of the university have also received very large research funds (e.g., VIDO and the College of Agriculture). This has resulted in USask having the largest per cent increase in overall research funding relative to any other major university in Canada (as shown in tables 3 and 7). To show this clearly, I have used the research funding (R) for 1999 and 2016, and taken the R2016/R1999 ratio to express the relative increase in funding. Table 7 shows that USask, with a ratio of 4.23 (shown in table 3), has the largest increase in research funding, while my home university of Western (a much larger institution than USask) is near the bottom of the chart, at number 15. It is important to emphasize two other facts as well. First, even without the more than $30 million a year in CLS funding, USask would still be in the top seven, with

118  The Canadian Light Source

a ratio of 3.6. Second, Western’s research funding has decreased greatly relative to the more comparable universities of McMaster, Calgary, Laval, and Ottawa. Table 3 shows that these four universities (with similar research funding in 1999) had at least $100 million more in research funding than ­Western in 2016. Specifically, McMaster had $9 million less in r­ esearch funding than ­Western in 1999, but it had $120 million more in 2016. As someone who has spent most of his career at Western, and who has tried very hard to ­elevate its international reputation, I find these poor Western numbers e­ xtremely worrying for the future of my home university. I would like to spend a few paragraphs describing some of the reasons for these two very different research trends at USask and Western. In his book,12 Peter MacKinnon describes some of his efforts to improve the research performance at USask after he was appointed president in 1999. As background, Walter Murray, the first president of USask from 1908 to 1937, said that “this is a university of the people, established by the people, and ­devoted by the people to the advancement of learning and promotion of happiness and virtue.” Murray also wanted USask to have “an honourable place among the best.” Such sentiments reflect the problem that arose during the presidential selection process in 1994 at Western (described in chapter 4, section bii). Was Western (or USask) going to concentrate mainly on its undergraduate role as a liberal arts/science institution with some strong research units? Or was Western (or USask) going to try to be an outstanding research-intensive university and take its place “among the best”? MacKinnon tried to shift the university’s culture early in his tenure.12 He emphasized “excellence based on competition,” arguing that “excellent research does not compromise the undergraduate role of a university.” Both of those ­assertions have been self-evident to me for my whole career; but it was not evident to a significant fraction of the Canadian university community twenty years ago. For example, many faculty claimed that USask was unique with its emphasis on local service, and that its natural role was as the “people’s university.”10,12 These faculty thought (as many still do) that a concentration on research (especially with industrial connections) would compromise this status. In contrast, President MacKinnon asked in the 2000–1 annual report: “Is it possible for a province with a small population and limited tax base to support a world-class educational and research institution”? It should be noted that this emphasis on competition and excellence came from a lawyer; and lawyers at universities are generally much less concerned with competitive research and research grants than faculty in science, engineering, or medicine. MacKinnon began an extensive integrated planning exercise to identify ­areas of strength in the university that could be selectively enhanced to a truly ­excellent national or international level, sometimes called targeted investments in research. The CFI promoted “excellence based on competition” in a big way, and most universities set up committees to identify strong research units for

The Positive Impact on USask and Canadian Science  119

“selective enhancement.” Such planning inevitably is very controversial, b ­ ecause most faculty think that their department should be “enhanced” for one reason or the other. Also, as demonstrated by the reaction I received from USask faculty in 1999, there is always the suspicion that any enhancement of one department or area leads to a loss of funding for another. Michael Hayden wrote in 2002 that no planning was needed, and that planning exercises had been incredibly rare at USask because of the vision of the founders. Debate was so heated in 2001–2 that a number of faculty established the People’s Free University to oppose the plan of enhancing excellence and competition. Indeed, the faculty union greeted MacKinnon’s plan not only as inappropriate but as illegal!12 It was apparent to most at USask by 2001 that CLS was not syphoning money out of other departments – on the contrary, all indications were that it would enhance many of them (see tables 5 and 6) by providing unique and excellent opportunities for many faculty across much of the university. I think CLS probably made it easier to sell the idea of selective enhancement in the strategic directions statement presented to the university and its Board of Governors on ­November 24, 2001. Despite the intense internal wrangling, MacKinnon pushed ahead with his planning exercise. For any such planning to gain broad support (and to succeed), the enhanced area must be of the highest quality, and this quality is traditionally based on research funding, research productivity, and impact (often measured by numbers of publications and citations). USask did a good job in selecting a few very specific areas for enhancement (very difficult to do), whereas Western took the politically expedient (and much easier) route of selecting many areas of the university as excellent. This is one reason for vastly different increases in research funding for the two universities over the 1999–2016 period. After many years of discussion (unsuccessful to begin with, and probably very acrimonious) the vice-president research of USask, Karen Chad, led the process to identify just six areas that were approved: (1) Indigenous peoples; (2) agriculture; (3) energy and mineral resources; (4) one health; (5) synchrotron sciences; and (6) water security. All of these areas have been enhanced with a combination of CRCs and CERCs (Canada Excellence Research Chairs), many other research chairs (118 in all), and distinguished professorships. Major funding came from CFI, greatly enhanced with WED and provincial monies, and in some cases industrial support. I have already mentioned the large amount of CFI funding to CLS and VIDO. But there are several other examples of multi-million-dollar grants in agriculture and water security. For example, USask was successful in the first CERC competition, which gave them two out of thirty chairs across Canada in food security and water resources. In contrast, Western only received one CERC for the Brain and Mind Institute, the most prominent area of Western enhancement. While the difference in academic concentration of the two universities is large, and the situation is complicated, there can be little doubt from tables 3, 7, and 8 that USask has done a very good job in enhancing research, while

120  The Canadian Light Source Table 8.  Increases in research funding from 1999 to 2016: as given by the ratio of the funding in the two years in table 3 University

Rank

University

Rank

Saskatchewan McMaster Laval Sherbrooke British Columbia Ottawa Calgary Toronto Dalhousie

1 2 3 4 5 6 7 8 9

Waterloo Manitoba McGill Montreal Alberta Western Queen’s Quebec Guelph

10 11 12 13 14 15 16 17 18

Western has done one of the poorest jobs of any such institution in Canada. From my first-hand experience with both universities over many years, there is no question in my mind that the difference in trends at the two universities is due mostly to leadership. MacKinnon muses in his book whether a president can make a real difference to a university. I believe that the vision of Presidents Ivany and MacKinnon is the major reason for the improvement in the research funding and image of USask since 1999. They believed that research excellence was critical. In contrast, the Western administration believed that enhancing the undergraduate experience was the top priority, with research excellence being of secondary importance. The undergraduate experience at Western has indeed been greatly enhanced, and Western now has a very high popularity among undergraduates. The incoming average of entering students at Western went up dramatically, and Western has often been ranked first in Canada for “the best undergraduate student experience.” More importantly for me, Western has always been a great place to work. But the lack of focus on research excellence (as evidenced by the lack of response to the CLS project at Western) has led to a poor research performance relative to Western’s competitors and has resulted in a large slip in its international rankings (from its previous place far below the 200th position to greater than 200th), such as in the QS World University Rankings and the Times Higher Education Rankings, since the 1990s. Just recently (September 2018), Western’s ranking in one survey rose to the 190th position, after a lot of effort by President Chakma and senior administration at Western to nominate outstanding researchers for Canadian and international awards. In contrast, McMaster’s rankings have greatly improved, rising to 75th position. Somewhat surprisingly, USask’s international rankings have not yet obviously improved, and the university is not in the top 400. But that should change significantly in the next few years with the USask research at CLS, and the several outstanding hires in other areas. However, I believe that USask would have benefited

The Positive Impact on USask and Canadian Science  121

much more from research at CLS, and its rankings would have improved, if it had made two significant decisions. First, the many talented CLS scientific staff should have been integrated into USask departments through adjunct appointments; and second, a strong synchrotron institute should have been formed at USask. Had either decision been taken, many more papers and citations would have accrued to USask, because USask and/or the USask institute would appear on many more papers. c) The Impact of the Initial CLS Funding and Philosophy on the CFI and Canadian Science As mentioned in the previous chapter, the CLS project was at least partly responsible for the formation of the CFI, and was important for the conceptual and philosophical basis for the entire CFI program. The great effort from the Saskatchewan group and the academic community (through CISR) to secure funding for CLS from 1995 to 1997 certainly influenced Paul Martin and the federal government to form the CFI. This effort to identify CLS funding led to the initial CFI allotment, and eventually to the $7 billion in total CFI funding to Canadian universities and hospitals from 1999 to 2018. Also, CLS helped the CFI promote national projects, since CLS had strong written support from more Canadian universities (eighteen) and industries (forty) than any other Canadian project in history. CLS was by far the largest facility funded in the initial CFI competition, and remains the largest CFI contribution to a university to this day. The CLS concept of a national facility on a university campus was important for defining the major goals of the CFI: to support and enhance world-class fundamental research and technology development in Canada by investing in state-of-the-art facilities and equipment in universities, colleges, research hospitals, and non-profit research institutions. The CFI’s initial objectives were to promote interdisciplinary national and international collaborations. The CFI now uses the term “convergence,” which goes beyond interdisciplinary research by bringing many research fields together. The successful national CLS synchrotron facility has to be the Canadian “poster child” for promoting convergent collaborations of all types. I know of no other scientific facility in Canada that supports so many researchers covering so many scientific disciplines (as an example, see table 5 for USask involvement from 27 different departments) while delivering strong technical and scientific help to optimize results. CLS is now close to providing 25 separate independent beamlines (all with different characteristics) with close to 100 different experimental chambers, which give unique chemical and structural information on every type of matter. No single laboratory in Canada can match the quality of results from any of those beamlines. Moreover, CLS facilities are relatively easy to access. For example,

122  The Canadian Light Source

academic and industrial protein crystallographers can mail their samples to CLS, and a CLS crystallographer or industrial scientist will obtain the data for the users, often with the interaction of the researcher in a remote mode. In most other cases, the researchers need to travel to Saskatoon, often with the aid of student travel grants from CLS. Usually with advice from a long-term CLS user or staff from the beam team, a new researcher can access CLS with a small application, which is then peer-reviewed. If approved by peer review, further help is then provided, either by the beam team or the CLS beamline staff to plan the experiment in detail. A nearby residence, owned by CLS, is available for accommodation at relatively low cost. Co-operation and often collaboration from the beamline staff is essential to optimize results and to attract the researcher back for additional experiments. Indeed, CLS has become essential for a very large number of Canadian and international scientists in many disciplines. (Some examples will be shown in the next section.) But in addition to collaboration, international co-operation and expertise is, and has been, critical at CSRF in Madison and at CLS in Saskatoon. Examples have already been given for CSRF. The director of SRC, Ed Rowe, was an outstanding example of international generosity. He welcomed and encouraged Canadians to establish CSRF at a foreign laboratory hundreds of miles from Canada. Several other Americans at the SRC, such as Charlie Pruett and Hartmut Hochst, helped CSRF with beamline and experimental developments. I have already mentioned the international staff directly employed in establishing the first two beamlines at CSRF; as we have seen, Kim Tan from Malaysia, T.K. Sham from Hong Kong, Masoud Kasrai from Iran, Jin-Ming Chen from Taiwan, and B.X. Yang from China and the United States were critical to the Canadian synchrotron effort, which would have been delayed were it not for these individuals’ efforts. There has been a great effort at most Canadian universities to promote ­international education and international co-operation; and the ethnic mix of students and staff at Canadian universities has changed dramatically in the last ten years. CLS demonstrates the great value of international staff, students, post-docs, and faculty. On the technical side, Les Dallin’s accelerator designs were checked by several international accelerator designers; and international beamline experts often reviewed beamline plans and helped out with beamline problems. There are at least three important reasons why this international collaboration should be greatly promoted. First, as noted above, for CSRF and CLS, such co-operation is often essential to optimize equipment and research r­ esults for fundamental and applied research. Second, this collaboration is also necessary for the economic benefit of Canada, although the time frame over which such benefit occurs may be many years after the research (as seen b ­ elow). Thomas Friedman, the well-known New York Times columnist, expressed well the ­importance of international collaboration in a May 27, 2007, article for that

The Positive Impact on USask and Canadian Science  123

paper: “It is pure idiocy that [the US] Congress will not open our borders as wide as possible to attract and keep the world’s first-round draft choices in an age when everyone increasingly has the same innovation tools, and the key differentiator is human talent.” Since the election of Donald Trump in 2016, the United States has shut out foreign talent much more, with very positive effects for Canada. For example, many Iranian students and post-docs are flocking to Canadian universities, partly because they are not welcome in the United States. As of 2018, the chemistry department at Western had close to 40 international students, with 18 coming from China and 12 from Iran. These students are very talented, well-trained, co-operative, and motivated; and they have had a very positive effect on the research at Western and CLS. There is, indeed, competition for the best talent; and with a lot of international talent now shunned by the United States, Canada is reaping the benefits. Third, and I believe most importantly, it is critical for the future of humanity for all people of good will to meet, work together, and exchange scientific ideas, along with cultural and religious differences. It seems obvious to me that we can only deal effectively with the major problems of the world, such as climate change, viruses, racism, and nuclear proliferation, with global co-operation and collaboration. But the rise of populism in many Western countries – including the United States, Great Britain, and much of Europe – threatens humanity’s ability to address many of these important problems. Canada has been accepting many immigrants in the last few years, and has had considerable success in integrating many different ethnic groups. It is also encouraging that Maxime Bernier and his People’s Party of Canada, which campaigned on decreasing the number of immigrants, did not win a single seat in the 2019 federal election. We are, after all, a nation of immigrants. It is worth quoting here a paragraph from a September 12, 2018 article by Andrew Coyne in the National Post: “Certainly it’s valid to worry whether people from different backgrounds can form a coherent whole, able to work on common projects, aspire to common ideals, or at the least avoid conflict. But the evidence is that greater diversity helps, rather than hurts, in this regard.” Indeed, all the evidence from my career shows that international co-operation greatly helps teamwork and productivity. The vast majority of people of all races and ethnicities share common values, such as honesty, dedication, hard work, respect and tolerance, empathy and generosity. Moreover, most immigrants are very clever, well-motivated, extremely hard-working, and excellent team players. I want to finish this section with the most striking example I can provide of the important role that immigrants have played in my professional life. As mentioned in chapter 3, Masoud Kasrai, who conducted research for twenty years at CSRF in Madison (commuting from London, Ontario), was a graduate student at the University of Cambridge, and worked across the bench from me

124  The Canadian Light Source

in the Chemistry department from 1964 to 1968 in Dr. Alfie Maddock’s group. He was a disciplined liberal Shiite Muslim who prayed at least three times a day. I was a liberal Protestant Christian who went to Great St. Mary’s Church in Cambridge Sunday evenings with four British friends from my residence (Gavin Currie, Gordon Foster, Richard Haworth, and Ian Stenhouse), and, after 1967, with my new wife Joan. We heard outstanding sermons from people such as the archbishop of Canterbury and the British prime minister. I had never met a Muslim when I went to Cambridge, but I immediately respected Masoud and enjoyed his company. We had tea almost every day for close to three years; and we shared many things in common, including hard work, generosity, the best of our common religious teachings, and a lot of high-spirited humour. Although we did not socialize a lot after work, we shared an excellent Easter holiday in France with another friend, Paul Davies, in 1965. I learned a good deal about Islam and Iran from Masoud; and the many positive human values common to both Islam and Christianity. For example, both religions (and ­indeed all major religions and every ethical tradition) share the importance of the Golden Rule first enunciated by Buddha and Confucius around 500 BC. The Christian version is “Do unto others as you would have them do unto you.” Dr. M ­ addock’s group of ten students included eight international students, and I learned from Masoud and all of the group the most important lesson of my life: that all people have great value and possess unique talent, irrespective of race, colour, creed, class, sex, or sexual orientation. Masoud went back to an academic position in Tehran in 1968. Because of the Iran-Iraq War in the 1980s, his family wanted to leave Iran and work in North America. I was able to arrange a senior CCP fellowship at Western. Masoud, his wife Guiti, and his two teenage children, Leila and Reza, came to Canada in January 1987. He drove back and forth to CSRF in Madison many times a year for close to twenty years, often “chauffeuring” graduate students and post-docs. He helped many of my graduate students with their results and the interpretation of those results, including a number of students in the tribology area (see page 36). One of my graduate students, a fundamentalist Jew, was a strong supporter of Israel and its usual policies toward the Palestinians. Also, initially, this student was not particularly motivated and needed extra supervision. Because of the strong antagonism between Iran and Israel, the vast majority of Muslims would not be comfortable with an outspoken, right-wing supporter of Israel. Surprisingly to me, Masoud not only tolerated the student, but helped him a great deal in a number of ways. He ended up writing several good papers and a very good PhD thesis, and he and Masoud finished as good friends, despite the large differences in their backgrounds and beliefs. I should also mention that Masoud’s wife Guiti has her PhD in nuclear physics from University College, London, and worked successfully in research and

The Positive Impact on USask and Canadian Science  125

administration at Western for over twenty years. Masoud and Guiti’s two children were extremely bright; they have contributed greatly to Canada, and indeed other countries as well. Leila is a plastic surgeon in the medical faculty of the University of Toronto, and Reza (with a PhD from McGill in medical imaging) has devoted the last fifteen-plus years to helping the poor in developing countries, usually through his work with non-profit aid groups. The children of all my other international students are similarly well-educated, and are making excellent contributions in Canada and abroad. Scientific scholarship, and indeed academic scholarship in general, is an excellent way of enhancing international co-operation and fostering strong relations between all human groups. The present Canadian school/university setting, with a strong representation from all major ethnic groups, makes it much more difficult to discriminate when no group is in the majority; and all indications are that school/university students of all ethnicities co-operate very well. Scientific research is inherently very competitive, and such competition can result in personal conflicts within or between research groups, Canadian or international. But from my experience, it is quite unusual that two contemporary research studies overlap greatly, mostly because research in many areas is virtually infinite. Also, all global scholarship is now available to all researchers through the Internet, and all groups use other researchers’ publications as background for their studies. The international synchrotron community co-operates very well because synchrotron science is so immense and because each facility contributes a unique mix of research. As a result, every good facility has a lot of international users, who come for some of the unique areas that the facility is known for. Such close co-operation and collaboration is not as usual in industry because of intense competition, intellectual property concerns, and the profit motive. d) The Importance of CLS Research-Convergence and Internationalization In the 2018 annual report, CLS’s director, Rob Lamb, states that “scientists from 30 universities in seven Canadian provinces as well as 15 countries published more than 500 scientific articles.” These articles span a very large number of scientific areas, as indicated by the twenty-seven departments at USask that use CLS (see table 5). The breadth and depth of the research at CLS is unequalled by any single Canadian research laboratory, and it seems important to describe this research generally before discussing a few outstanding papers. There are four major strategic research areas, as outlined in Rob Lamb’s ­message in the 2018 annual report. The largest strategic area, that of advanced materials, includes research in many disciplines and sectors, including new insights and applications in high-temperature superconductors, fuel cells,

126  The Canadian Light Source

batteries, eco-composite materials, solar power, catalysts, micro devices, and other new materials such as nanotubes. The health area includes diagnosis and drug development related to many diseases, including cancers, heart disease, cystic fibrosis, and antibiotic-resistant infections. Much of this research relies on high-resolution crystal structures of complex proteins obtained on CLS x-ray diffraction beamlines. Over a thousand such structures have been solved at CLS and deposited in the international Protein Data Bank. The environmental area, which includes industrial contract work for Saskatchewan-based companies, includes advancement in mine remediation techniques and remediation of contaminated ground water and soil, heavy oil extraction efficiencies, highly efficient catalysts for petroleum refinement, renewable resources, and energy storage. The agriculture area is critically important for the Saskatchewan and Canadian economies. This area of research, performed on several beamlines, is not actively pursued at other international synchrotron facilities. It includes new insights into improved crop development, fertilizers, drought and temperature resistance, and soil management. It is very difficult to select a few papers to highlight from the over fifteen hundred published since 2005. I have used five criteria for my selection. First, I want to describe at least one paper from the four main strategic research areas outlined above. Second, I want to highlight convergent research contributions with many international researchers from several different departments and universities. Third, I want to show that many of Canada’s top researchers in diverse areas are taking advantage of CLS for their research. Fourth, I want to show that many of the papers from CLS are in very prestigious journals (e.g., Science, Nature, Physical Review Letters) and are very highly cited. And fifth, I want to highlight research on most of the working CLS beamlines. ­Because most ­ research today requires the best analytical techniques, CLS makes the possibility of a groundbreaking discovery and another Nobel Prize to a ­Canadian scientist more likely in the near future. Of course, ideas usually come from ­individuals, so many important research discoveries come from smaller groups; and strong leadership is essential for large groups. Some of this section is influenced by highlights from recent CLS annual reports. i)  Advanced Materials Thousands of scientists worldwide are currently working: to improve existing battery technology; to understand the best present high-temperature superconductors so that superconductors, operating at room temperature, can eventually be discovered; and to improve catalysts for many important chemical reactions such as splitting of water into hydrogen and oxygen. In the above areas, outstanding Canadian scientists, supported by large interdisciplinary international

The Positive Impact on USask and Canadian Science  127

teams, have used CLS extensively in the last few years. Their research has been published in the very best journals, such as Science and Nature, and has generated much international attention. Most importantly, CLS has been critical to this important research. Dramatic improvements to current batteries and high-temperature superconductors would probably result in a Nobel Prize and lead to huge industrial spin-offs. CLS makes it much more likely that such ­developments will occur in Canada. In the following sections, I want to summarize briefly CLS-based research in the above three research areas. 1)  batteries A greener future is critical for the human race; and a greener future depends, in part, on better batteries for storage and electric vehicles. Thousands of ­researchers worldwide are working on improving existing lithium batteries for electric vehicles, with the hope that these batteries will be smaller, cheaper, safer, and have a range of over a thousand kilometres on one charge. At Western, a large group of about forty researchers from the Department of Mechanical and Materials Engineering (with others from the Departments of Chemistry and Physics) and originating from Canada, Iran, China, and ­Russia, is led by Andy Sun in the Department of Mechanical and Materials Engineering. They have been studying many different types of lithium and sodium batteries for the last ten years. Sun joined Western over fifteen years ago and quickly set up his group – Advanced Materials for Clean Energy – to study both battery technology and fuel cells. T.K. Sham introduced Sun to the power of CLS about ten years ago; and Sham also introduced many others to CLS and other synchrotrons through his formation and leadership of the Chinese Soochow ­University-Western ­University Centre for Synchrotron Radiation Research, begun in 2012. Use of the soft x-ray microcharacterization beamline at CLS has greatly enhanced the understanding of battery reactions, and it has produced many highly cited papers in journals such as Science and Nature, as well as local and national news coverage. Indeed, several of Sun’s recent papers (over 180 since 2015) are listed as highly cited and/or “hot” research papers, and the Web of Science website has named him a Highly Cited Researcher. This list recognizes world-class researchers for their exceptional research performance, demonstrated by production of multiple highly cited papers that rank in the top 1 per cent by citations for field and year. Many of Dr. Sun’s group now travel to CLS regularly, and some of his staff (e.g., Mohammed Banis) have spent up to a year at CLS developing new techniques with CLS staff. Dr. Sun’s research has involved a large interdisciplinary and international team that includes the beamline staff scientist Yongfeng Hu, as well as Mohammed Banis. Banis has noted the critical importance of the large team. After publication of a recent

128  The Canadian Light Source

paper, he told me, “without all of them, I don’t think we would have gotten to do this research. But if we bring people from different places together we can achieve something remarkable.” Sun and his team have recently spent considerable effort at CLS on characterizing the chemical reactions involved in lithium-sulphur batteries. This battery is one of the most attractive energy storage systems because of its better theoretical performance compared to today’s lithium-ion batteries.35 But there have been considerable safety and stability problems with ­lithium-sulphur ­ batteries, as outlined in reference 35. With their research in the last few years, Sun and his team have circumvented some of those problems. Dr. Banis was responsible for constructing a small battery cell with a very thin window to let low-energy x-rays in and out that could be used in a CLS beamline to follow the reactions of the sulphur in situ during charging and discharging. For the first time in a working lithium-sulphur battery, the different chemical types of sulphur (e.g., elemental S, sulphide, polysulphide, and sulphate) could be readily distinguished from the spectra. Most importantly, this battery arrangement did not produce polysulphides, which damages battery performance and lifetime. In a January 2019 CLS web release, Dr. Sun writes: “Close collaboration with CLS to obtain such detailed information is very important to our understanding of batteries. We need not only design novel materials for ­energy storage, but also deepen understanding in the science behind the materials.” A battery R&D company, GLABAT Solid State Battery Inc., has been created at the Western research park to test different battery designs as well as marketable new batteries. Sun and the new company have received over $13 million in funding from the China Automotive Battery Research Institute Co. Ltd. Sun told me recently that Western should be number one in five years, “not only for fundamental research innovation, but also for the fabrication of solid-state batteries in Canada.” 2)  superconductors Present superconductors are critical in many scientific areas, including magnetic resonance imaging (MRI) and in the CLS synchrotron ring. Most of these become superconducting below 10 Kelvin (K) (−263 centigrade (C)), requiring liquid helium to achieve superconductivity. The first high-temperature superconductor was discovered in 1986 by Georg Bednorz and Alex Muller from IBM Zurich Research Laboratory. They were awarded the 1987 Nobel Prize in Physics for their important breakthrough in the discovery of superconductivity in ceramic materials such as barium-doped compounds of lanthanum and copper oxide that became superconducting at about 90 K (the transition temperature). Up to 2015, the highest transition temperature was 135 K, making superconductivity available with liquid nitrogen at 80 K. If room temperature superconductors become viable, they could be used to

The Positive Impact on USask and Canadian Science  129

create ultra-efficient power grids, supercomputers, and magnetically levitating vehicles. ­Canada has one of the most outstanding groups of physicists in the world studying superconductors, headed by the UBC physicists George Sawatzky (one of the most-cited physicists in the world) and Andrei Domascelli, and David ­Hawthorn at the ­University of Waterloo. The UBC scientists in the physics and astronomy d ­ epartment and the Quantum Matter Institute have also collaborated with the Toronto physicists from the Quantum Materials Program at the Canadian Institute for Advanced Research. These Canadians, together with CLS staff scientists Feizhou He and Ronny Sutarto and an ethnically diverse group of over twenty physicists from China, France, Germany, and Italy, have used CLS resonant inelastic scattering beamline and other similar beamlines at international synchrotrons to understand how superconductivity occurs in a range of complex copper oxides. Their early papers in 2012 in Science and ­Physical Review Letters have been widely cited, with close to six hundred citations in one case.36 More recent papers in Science37 and Nature have also been widely cited. These studies contribute to the still elusive theory of ­high-temperature superconductivity, which should result in the preparation of room temperature superconductors in the near future. 3)  catalysts Many groups have studied catalysts and catalytic reactions using a few of the CLS beamlines. Catalysts are used to dramatically increase the rate of important reactions such as hydrogen and oxygen evolution from water, conversion of the greenhouse gas carbon dioxide to ethylene (used for plastics), and ­removal of sulphur and nitrogen from bitumen during its conversion to synthetic heavy oil. In all of these studies, the chemistry of the catalyst, obtained from the synchrotron studies, is important for understanding and improving on the catalysts for more rapid controlled reactions. These studies have resulted in many highly cited papers in high-impact journals. The oxygen reduction ­paper published in 2011 by H. Dai and colleagues38 now has over three thousand citations, making it the most highly cited paper to emerge from research done at CLS. This work was a collaboration between a research group led by H. Dai in the Department of Chemistry at Stanford University and CLS scientists and beamline staff scientists Jigang Zhou, J. Wang, and Tom Regier using the SGM beamline at CLS. A very large group of about seventy researchers from many countries, headed by E.H. Sargent from the University of Toronto’s Departments of Electrical and Computer Engineering and Materials Science Engineering, have produced efficient catalysts for splitting water39 and conversion of CO2 to valueadded chemicals such as ethylene. They are routinely using three of the CLS beamlines for chemical characterization of these catalysts, and are aided by beamline staff scientists Ning Chen, Tom Regier, and Yongfeng Hu.

130  The Canadian Light Source

ii) Health 1)  protein structures Proteins are the very large organic molecules containing one or more chains of amino acids that are an essential part of all living organisms. They are critical to build, maintain, and repair all body tissues. There are four general types of proteins: molecules such as haemoglobin, which transport other chemicals (e.g., oxygen) in the body; enzymes such as sucrase, which catalyze chemical reactions; antibodies such as blood plasma proteins, which ward off disease; and hormones such as insulin, which regulate body functions. There are about 20,000 human genes capable of producing proteins, and each gene can produce multiple forms of a protein. The number of proteins in the human body is hard to estimate, but there could be as many as 100,000 different proteins in one cell. Proteins are extremely large molecules with molecular weights from about 10,000 to more than 300,000 Daltons. Their structures are extremely complex, and it is necessary to obtain the atomic structure of these large molecules to understand protein ­functions and to design molecules (drugs) that can enhance or retard these functions. To obtain publishable protein structures today, a laboratory x-ray source is seldom sufficient; rather, a third-generation synchrotron is essential to solve the structures at the atomic level, especially on very small crystals with very large m ­ olecular weights. Indeed, about 90 per cent of the solved protein crystal structures are coming from the over 100 protein crystallography beamlines worldwide. The first protein structure (of myoglobin) was obtained by Max Perutz and John Kendrew with great difficulty in the late 1950s, and for their efforts they won the Nobel Prize in Chemistry in 1962. Even in the 1960s and ’70s, very few structures were solved with laboratory sources and limited computing ­facilities, and yet several of these structures also resulted in Nobel Prizes. The advent of many intense, focused synchrotron beamlines caused an explosion of protein r­ esearch and new protein structures. As of 2018, over 140,000 structures of protein and protein/drug combinations have been solved and deposited in the P ­ rotein Data Bank, with over 1,000 coming from academic and industrial ­researchers at CLS since 2005. Brian Kobilka, from Duke and Stanford Universities, won a ­Nobel Prize in Chemistry in 2012 for his synchrotron-based structural/­function studies on plasma membrane proteins, termed “G-protein-coupled receptors.”* These proteins are extremely important for processing information in many human cells, and also for the senses of sight, smell, and taste. They are extremely hard to crystallize, and the smallest s­ ynchrotron beams were critical to obtain atomic structures on the extremely small crystals that were eventually * Kobilka’s Nobel Prize lecture is available at https://www.nobelprize.org/prizes/chemistry/2012/ kobilka/biographical/.

The Positive Impact on USask and Canadian Science  131

produced. These structures showed how these proteins communicate in cells; and they should lead to future drug discoveries. The Canadian Macromolecular Crystallography Facility40 at CLS is composed of two beamlines and serves more than sixty-five Canadian and some international laboratories. Pawel Grochulski and many co-workers at CLS ­developed both beamlines, which are competitive with other international beamlines. ­Local users collect data directly at the beamlines, whereas the ­majority of other users collect data remotely or use mail-in service, with the data collected by staff members. With over a thousand structures obtained at CLS, it is not possible to do justice to the breadth and importance of this research, but a couple of examples hopefully will be informative for the non-specialist reader. One of the first highly cited papers from CLS41 came from a collaboration of University of Calgary scientists, CLS scientists, and others from Spain and the United States. This paper describes the study of proteins involved in the Norwalk virus, which is the cause of gastroenteritis outbreaks in many countries. The x-ray structures show that the position of one of the nitro groups in the protein could aid in the design of new antiviral drugs. In a second study, another large international group (from France, Germany, Portugal, Germany, and the University of Manitoba)42 have used CLS, the APS in Chicago, and the Swiss Light Source in Villigen, Switzerland, to obtain the atomic structure of the protein netrin-1 bonded to receptors or adducts. The interaction of this protein with a receptor molecule (UNC5) is responsible for cell proliferation and cancer. It is important, then, to find another molecule (drug) to prevent the netrin-1/UNC5 interaction. A detailed crystal structure of netrin-1 shows the part of the netrin-1 structure where the interaction occurs. Using this information, the researchers designed an antibody to disrupt the interaction between netrin-1 and UNC5. They show that the antibody triggers the death of cancer cells under laboratory conditions. The antibody is now in clinical trials. 2)  iron and alzheimer’s disease The brain biochemistry associated with Alzheimer’s disease is a topic of great current research interest. It is well known that the disease is characterized by aggregations of the peptide fragment amyloid-beta (Aβ) in the brain. But the role of metal ions in the neurodegenerative damage is less well characterized. It has been proposed that metal ions such as ferric (Fe3+) and ferrous (Fe2+) are important contributors to the damage; and it is possible that the iron could be used as a diagnostic tool for Alzheimer’s disease. A large international group comprised of researchers from Keele and ­Warwick Universities in England, the University of Florida, McMaster University, and the CLS beamline scientist Jian Wang, have used the scanning transmission x-ray microscope beamline (STXM) pioneered by Adam Hitchcock and his group at McMaster to study the role of iron in Alzheimer’s disease.43

132  The Canadian Light Source

The group investigated the distribution and oxidation state of nano-scale iron in brain tissue taken from a mouse model that shows deposition of the Aβ peptide. A beam of x-rays from the CLS beamline is focused on a thin section of the mouse brain tissue, and the sample is scanned over a 50 × 50 μm (μm = micrometres) area with the microbeam to yield a two dimensional image of both the carbon in the peptide and two different types of iron. It was possible to detect iron deposits down to approximately 100 nm (nm = nanometres) in size. For perspective, the width of human hair is between 20 and 180 μm. They found a direct correlation of the amyloid plaque with the iron, and evidence for an iron-amyloid complex. These results could be important in the future for developing iron-modifying drugs to combat Alzheimer’s disease. 3)  lung imaging for cystic fibrosis It is important to highlight outstanding medical research from the two biomedical imaging and therapy beamlines at CLS, headed by Dean Chapman of the ­Department of Anatomy and Cell Biology at USask, with critical contributions from Juan Ionowski at the university’s faculty of medicine. Again, the research ­depends on collaboration and co-operation of a large cross-disciplinary group from ten different countries (see, for example, references 44 and 45). Many different imaging and radiation therapy experiments have been published. For example, ­synchrotron-based microbeams for cancer therapy is promising because the beamline can deliver intense beams of x-rays smaller than a human hair directly across a tumour using low dose rates, resulting in minimal damage to normal tissue. The recent research on cystic fibrosis, an incurable genetic disease, is ­especially important. Most of the morbidity and mortality from lung disease is thought to be caused by the failure to clear bacteria in the lungs, leading to chronic inflammation. A layer of airway surface liquid (ASL) helps to remove bacteria, but no one has tested whether the inhalation of bacteria triggers ASL secretion. Dean Chapman has been a world leader in developing new x-ray techniques to enhance contrast for soft tissue and liquids such as ASL. These techniques, such as diffraction-enhanced imaging and phase contrast imaging, are based on refraction or phase effects rather than absorption, as is normally used. The latter technique was used to measure the height of the ASL layer (from 16 to 400 micrometres – about the width of a human hair) after introduction of bacteria44,45 in normal swine from the USask Prairie Swine Centre. They suggest that in swine or patients with cystic fibrosis, the inhalation of b ­ acteria would fail to trigger ASL secretion, leading to infection and inflammation. iii) Environment and Agriculture Even before CLS began operation in 2005, the facility’s management decided to have mining and environmental studies as a major area of research, and I hired

The Positive Impact on USask and Canadian Science  133

Jeff Cutler in 2000 to begin developing this area of research with Canadian ­industry using US synchrotron facilities. Before operation of CLS, Jeff developed several strong clients in the mining area prepared to use CLS. Synchrotron radiation is especially important to the uranium mining companies, such as Cameco and AREVA (now Orano), in northern Saskatchewan. The initial synchrotron research conducted at US synchrotrons before CLS began operating in 2005 was immediately important for understanding the long-term chemical stability of toxic arsenic in the mines’ tailings ponds. About ten years later, Chitra Karunakaran at CLS began ­plant-innovation research of great potential importance to the large agriculture community at USask and the Western Canadian provinces. Improved crop and plant d ­ evelopment and fertilizer development are important components of that r­ esearch. Little agricultural research is done at other international synchrotrons, partly because most synchrotrons do not promote research on “dirty” samples such as mine wastes and soils, which are not immediately compatible with the high-vacuum conditions of the synchrotron and beamlines. ­Extra pumping capabilities and novel sample cells are required in many cases. B ­ eginning with our research at CSRF in Madison in 1990, sponsored by Jim Brown at Natural Resources Canada, we analyzed many “dirty” environmental samples such as coals and asphaltenes; and this type of research is a now a significant component of the work done at CLS. This attracts many international users who cannot do their research at a synchrotron near them. To show the practical importance to Saskatchewan of this more applied ­research, I want to give four examples in these two areas. 1)  preservation of organic matter in marine sediments The world’s largest carbon sinks for organic carbon (OC) matter (e.g., from forest fires) are the oceans and the soil. Understanding the mechanisms of OC uptake by minerals is critical to the preservation of OC in soil and sediments. Without this uptake, a lot of the OC will be oxidized to CO2 and released to the atmosphere, which in turn will increase global warming. In addition to clay minerals, it has been found that iron oxide nanoparticles are important for the removal of a large fraction of OC. The stable iron-OC complexes can persist for thousands of years in some sediments, implying that the OC is strongly bonded to the iron. The Fe-OC complexes have been studied previously, but most research has been conducted using model OC compounds and laboratory-prepared iron oxides that do not fully mimic natural conditions. The researchers used the STXM beamline (see example ii2 above) at CLS and a beamline at the NSLS in Brookhaven to study the iron-OC interactions at approximately 50 nm resolution in natural sediments collected from all over the world (e.g., the St. Lawrence estuary, the Arabian Sea, the Black

134  The Canadian Light Source

Sea, the equatorial Pacific Ocean, and seven others).46 The large interdisciplinary ­research group, headed by Y. Gélinas in the chemistry and biochemistry ­department at Concordia University, included oceanographic researchers from the University of Georgia and Marymount Manhattan College in New York. The most important part of the work used the STXM beamline at CLS with the involvement of the CLS beamline staff scientist J. Wang. The spectra showed that several C species were present, and that the OC was directly bonded to some of the iron in iron oxide nanoparticles. This has important consequences: the OC is stabilized and will not be converted to the greenhouse gases carbon dioxide or methane, and the reactive iron is also stabilized. 2)  stabilization of toxic arsenic in mine tailings Arsenic is a major toxic element in mining and smelting activities. The mining of high-grade uranium in northern Saskatchewan by companies such as AREVA and Cameco accounts for about one-third of the world’s total uranium production, a necessary component of nuclear reactors. Unfortunately, the processes to extract uranium from the chemically complex ore results in a large amount of soluble arsenic in “tailings” ponds. It is critical that the arsenic is “tied up” in very stable, insoluble compounds/minerals, so that arsenic is not released into the surrounding environment in the foreseeable future. To stabilize the arsenic, an iron compound is added to the arsenic solution to produce insoluble iron-arsenates. For licensing and operation, it is necessary for the companies to show strong evidence of stabilized arsenic to government regulators. The iron-arsenic species can be of three types, with one of them (the iron-arsenic mineral scorodite) being much more stable than the other two. Until the advent of synchrotron-based techniques, it was difficult to distinguish between these three types. But using the hard x-ray microanalysis beamline at CLS, the three different types of arsenic can be distinguished and ­quantified.47,48 The results, from over fifty samples from a tailings facility in northern Saskatchewan, showed that the majority species was usually the stable ­crystalline scorodite, with smaller amounts of the two less stable species. It was ­notable that crystalline scorodite was especially concentrated at the periphery of the tailings facility, where any arsenic could leach out more readily. The first study (see reference 47) has received many citations. The second very large study (see reference 48), involving CLS, McGill University, AREVA, and a consulting company in the United States, shows the important role that CLS played in helping industry understand the potential environmental impact of mining effluent. Moreover, it is important that AREVA has allowed this work to be published for use by other mining companies. A lot more research on arsenic speciation is ongoing at CLS by academics and other companies such as Cameco. For example, systematic studies of the effect of pH, temperature, and aging time have been undertaken to mimic the geochemical conditions at the tailings facilities.

The Positive Impact on USask and Canadian Science  135

3)  control of ammonia emissions from fertilizers by charcoal A very recent study by Cornell University scientists (see reference 49), combined with Australian scientists from Sydney and Adelaide and researchers from CLS, has shown that ammonia emitted from fertilizers can be adsorbed and stabilized by charcoal. This study, which was published in Nature ­Communications, involves several agriculture, soil, and plant scientists, along with spectroscopic expertise from CLS and the chemistry department of the ­University of New South Wales. CLS spectroscopic research on the phase 1 SGM, operated by the CLS beamline scientist Tom Regier, provided the most important information in this study. Indeed, one of the Cornell scientists is quoted in a CLS communication as saying that “the beamline capabilities at the Canadian Light Source were essential to this game-changing discovery, and turned it into a much bigger project than originally planned.” As an aside, it is important to note that Cornell University is much closer to the US facilities, such as Cornell’s own synchrotron CHESS and the NSLS, than it is to CLS. The Cornell researcher notes that “the unique endstations at CLS are great for this kind of nitrogen spectroscopy.” Ammonia, a common component of natural fertilizers, manures, and composting facilities, is critical for plant growth. However, ammonia is also a toxic gas that can damage human lungs. It contributes, moreover, to climate change when excess ammonia is converted to nitrous oxide, a potent greenhouse gas. In Canada and elsewhere, ammonia emissions have increased substantially since 1990, and global ammonia emissions should double by 2050. How can ammonia emissions be decreased? The earth’s soil contains large amounts of charcoal-like carbon, most of which originates from vegetation fires. This carbon is known to retain ammonia in soils, but it is not known how stable the ammonia-carbon interaction is. Using the CLS beamline facilities, which can distinguish different chemical types of nitrogen in any sample, the authors discovered for the first time that much of the nitrogen from ammonia is strongly bonded to the charcoal via several different types of carbon-nitrogen bonds. This discovery should have a dramatic impact on nitrogen balances in the soil and atmosphere after future research. 4)  clubroot resistance in canola plants Finally, I want to highlight research on one of the infrared beamlines.50 Most synchrotron beamlines optimize SR in the x-ray region; but all international synchrotrons have one or more infrared beamlines with substantial advantages over laboratory infrared sources. I also want to highlight at least one of the many recent plant studies at CLS, mostly resulting from the CLS scientist Chithra Karunakaran’s pioneering work with many different groups. Clubroot disease is a serious threat to canola production in western Canada, and indeed many parts of the world. With an annual value of $27 billion, canola

136  The Canadian Light Source

is Canada’s largest cash crop. Many varieties of canola are not resistant to the clubroot pathogen, or are losing their resistance. To make an important contribution to the mechanisms associated with clubroot disease, experts from CLS and several international scientists at the Saskatoon Research and Development Centre and Agriculture and Agri-Food Canada was essential. The present study 50 characterized the chemical changes in the cell wall of ­canola roots. Using the Fourier-transform infrared spectroscopy at CLS, chemicals such as lignin were found to be more prevalent in the roots that were ­resistant to the disease. They concluded that cell wall components such as lignin may well play a defence role against clubroot disease. Studies such as this will help us understand and ameliorate the disease. e) Industrial Impact and the Formation of a Spin-Off Company, Canadian Isotope Innovations Inc. CLS was committed to providing up to 25 per cent of the time on the beamlines for fee-for-service industrial contract work. From 2001 to 2005, the first industrial scientist, Jeff Cutler, attracted a number of Canadian industries for experiments at US synchrotrons in order to build commercial work for CLS when it opened in 2005. The industrial group, with about six staff scientists, has since concentrated on companies in the four major research areas listed above. The industrial effort led effectively by Jeff Cutler (and more recently Jeff Warner) generated close to $1 million annually between 2015 and 2018, with about half of that coming from the crystallography of proteins and protein adducts. The total industrial revenue of $4.1 million between 2005 and 2015 came from 139 industrial reports and 280 industrial projects. In 2017–18 alone, eighty-two technical reports were produced for 42 clients and 280 industry projects. And yet, the industrial research has never approached the 25 per cent target on most beamlines. However, compared to all international synchrotrons, the industrial activity and revenue at CLS is larger than at any other facility in the world, even if the revenue is still a small percentage of the overall annual operating budget of more than $30 million. Some academics at USask initially voiced great concern that the synchrotron would be dominated by industrial research and argued that universities should not be supporting such industrial efforts. Indeed, in his 2009 book, Howard Woodhouse from USask wrote that CLS is an example of the commercialization of university research that is endangering “other areas of scholarship and research that is perceived to be of little immediate utility for generating private monetary wealth.”51 The above numbers show that there is no danger of an ­industrial takeover of CLS! Moreover, it is important to emphasize that this

The Positive Impact on USask and Canadian Science  137

work does not significantly affect the academics using the facility, and that it is of significant benefit to many industries and the economy. As indicated above, protein crystallographers (both academic and industrial) have to use synchrotron sources to obtain high-resolution structures at the atomic level, especially on small crystals. Other industries need CLS to solve other environmental and materials problems. Much of the industrial work, such as protein crystal structures for the pharmaceutical industries, is run confidentially in a fee-for-service mode and not published in the open ­literature. The industries hold the intellectual property on the results, and CLS is often not aware of the impact or importance of these results. In other cases, CLS has worked with companies in a long-term contract mode as well as fee-for-service. For example, as noted above, CLS has worked with some of the uranium companies in northern Saskatchewan to help them understand the long-term chemical stability of their mine tailings.47,48 CLS has had a large economic impact both in Saskatoon and across Canada. A 2010 economic impact report included on the CLS website states that “CLS operations directly contributed almost $90M to the Canadian GDP.” This means that for every dollar of CLS operating funding (approximately $23 million in 2010), the CLS operation contributed almost four times that amount to the Canadian economy. With a much larger budget now, the $90 million impact in 2010 is much closer to a $150 million impact in 2018. Of greater industrial importance, CLS has formed an important spin-off company, Canadian Isotope Innovations Inc.,52 to produce radioisotopes for diagnostic medicine. Over ten thousand hospitals worldwide use these isotopes – one example being the radioactive isotope of technetium (99mTc) – for diagnosing many ailments; and the sales of these isotopes approaches $5  ­billion. Canada has been a major supplier of radioisotopes, using the ­fifty-year-old n ­ uclear reactor at Chalk River Nuclear Laboratories for decades. But that r­ eactor was due to be closed in about 2015 (it was finally shut down in 2018). In anticipation of a shortage of radioisotopes, in 2012 the Canadian government called for proposals for novel methods to replace the Chalk River ­production. Mark de Jong, the machine director at CLS, proposed that high-­ energy linear electron accelerators, similar to the 1964 SAL linear accelerator, could be used to provide a safe and reliable source of radioisotopes. The CLS Board chair, Walter Davidson, Director Josef Hormes, and the CLS Board were very supportive. Focusing on production of the very useful 99mTc, Mark applied for about $10 million for an electron accelerator and ancillary equipment. In this process, the high-energy electrons in a high vacuum impinge on a metal such as tungsten (W) producing very high-energy x-rays. The x-rays impact a small molybdenum (Mo) metal target of one of the isotopes of Mo (100Mo). The reaction of the x-rays with 100Mo causes one neutron to be released from the 100Mo to give another unstable isotope, 99Mo. This decays to 99Tc, the isotope of choice

138  The Canadian Light Source

for positron emission tomography (PET) in nuclear medicine. By 2015, Mark and his team showed that this method was a safe and efficient method for 99Mo production. The new director, Rob Lamb, along with the CLS Board, proposed that CLS would financially support a new spin-off company (CII) located in the research park. The really big problem has been raising the money required for a large-scale industrial facility, including several linear accelerators and ­efficient chemical separators for retrieval of the 99Tc from the 99Mo. The story of this development may well be more complicated than the story of CLS itself! ­Fortunately, the company is now headed by the former excellent CLS e­ mployees Mark de Jong and Beryl Lepage. It is nonetheless very difficult to get a major investor in Canada to bring this company to its full potential. Hopefully, CII will be a major world supplier of 99Tc and other isotopes in the near future. f)  Broad Educational Impact After the enormous football-field-sized building was constructed in 2001, S­ andra Ribeiro of CLS began offering regular hour-long tours for the public to see the facility and receive overviews of the expected science. After beginning operation, CLS expanded its educational programs for teachers, students, and laypeople. Tracy Walker and her group coordinate these programs, which are summarized below. I know of no other research laboratory in Canada that offers such educational programs. Indeed, these are an excellent model for science education and awareness in Canada. CLS is ideally suited to such tours because the mass of scientific equipment (accelerators and beamlines) in the huge 100 metre by 100 metre vault is about 6 metres below a mezzanine floor that extends for about 210 metres on three sides of the building. Large posters, which describe the facility and research, cover the mezzanine walls outside most of the CLS offices. Over 400 laypeople and students per month visited CLS in both 2017 and 2018 for these free, hour-long tours (conducted from 2:30 to 3:30 p.m. on ­Mondays, Wednesdays, and Fridays), and it is possible to book a tour online for five people or more on a different day or time. Over 50,000 people have visited CLS on these tours since the facility opened in 2005. For the tenth anniversary of CLS in 2015, about 1,200 people attended an “open house.” Three unique educational programs have been developed for primary students, secondary students, post-secondary students, teachers, professors, and educators: Light Source Student Experience, Students on the Beamlines, and Teacher Professional Development. In the first program, elementary and young high school students visit CLS and take part in physics experiments looking at properties of light. In one case, the monochromatization of light is demonstrated with the aid of a flashlight, a prism, and a piece of cardboard with slits cut into it. By rotating the prism, the students can show that the white light from the flashlight yields a single colour on the wall.

The Positive Impact on USask and Canadian Science  139

The second program has been funded for nine years by PromoScience, a program funded by NSERC. In this program, mentors help high school students use a CLS beamline to examine samples, usually geological or environmental, and to obtain new information of interest to the scientific community. The first step in attracting students to this program is attracting teachers to week-long summer workshops. This is an excellent professional-development opportunity for these teachers. Those that have attended such a workshop are eligible to create their team of students, and then work with CLS staff and other scientists at CLS to develop their own novel research program using a CLS beamline ­developed for general studies by Josef Hormes (CLS director from 2008 to 2014) and Kim Tan. The program is intended for small groups of student (3 to 15). Students from six provinces and the Northwest Territories have participated. In one case, they examined the effects of acid rain in northern Saskatchewan. In total, over 1,500 teachers and 20,000 students have participated in these programs, a terrific contribution to “hands-on” science education. Indeed, CLS was the winner of the Canadian Nuclear Foundation Education Award in 2012. At the university level, USask is offering a unique arts/science course in the College of Arts and Science, developed by Tom Ellis (of the USask chemistry department and formerly director of research at CLS), Tracene Harvey (of the classical, medieval, and renaissance studies department and the Museum of Antiquities), and Tracy Walker from CLS. In this seminar course, both arts and science students learn about spectroscopic methods for the study of the chemical and material composition and properties of objects such as medieval manuscripts and ancient glass, pottery, and coins. The class develops a plan to use one CLS beamline to analyze an ancient object from the museum, and then interpret and write up the results. At Western, Ron Martin from the chemistry department and Andrew Nelson from archaeology also give a joint course, Advanced Analytical Techniques in Bioarchaeology, where synchrotron techniques are introduced in two lectures. Several fourth-year undergraduate and PhD students supervised by Ron Martin, Andrew Nelson, and T.K. Sham have based their research on synchrotron techniques. Perhaps of greatest interest, a joint student, Madalena Kozachuk, has completed a PhD project using SR to recover the images from degraded old photographs (so-called Daguerreotypes).53 This paper has received worldwide attention in such media outlets as Voice of America, the CBC, and several US and Canadian newspapers. g)  The Future CLS has been operating for over fourteen years; and as we can see with SAL and the Aladdin Synchrotron in Madison, these facilities usually only remain competitive (and fundable) for twenty to thirty years. Over the last twenty years, in response to beamline and academic scientists who want more brilliant sources

140  The Canadian Light Source

of SR (that is, much smaller beams with much higher intensity), physicists have designed fourth-generation light sources to eventually replace the common third-generation rings like CLS. Fourth-generation rings must exceed the performance of third-generation rings by more than one order of magnitude in important parameters such as beam size, coherence, and number of photons in a given beam size. In a coherent beam, all photons have identical energies, phases, and amplitudes. These coherent beams are especially important for diffraction methods (so-called coherent diffraction imaging, or CDI), which often involve high-resolution imaging. There are two types of fourth-generation sources: a very large multi-bend achromat (MBA) storage ring, and an x-ray free electron laser (XFEL). The MBA storage rings utilize many more bending magnets (and other types of magnets) than third-generation rings to achieve the much smaller beams and higher intensities. MAX IV in Lund, Sweden, the first fourth-generation MBA ring, has been operating for over three years now;54 and at least ten other ­international fourth-generation MBA rings are being constructed or under study internationally. Eight beamlines at MAX IV are now accepting experimental proposals. This dramatic increase in performance increases the size and cost of the synchrotron facility greatly, to close to $1 billion. The other fourth-generation source, the equally expensive XFEL, is based on an extremely high-energy linear accelerator (LINAC). The linear coherent light sources at Stanford, the XFEL at Hamburg, and the SPring-8 Angstrom Compact free-electron laser in Japan have been working successfully for a few years. These sources can produce very short femtosecond (10−15 sec) coherent x-ray pulses that concentrate photons into much smaller (a few nanometres) and much brighter beams. Much larger pulse widths are produced by the MBA rings (several picoseconds, or 10−12 sec), although MAX IV produces femtosecond pulses from the 3 GeV linear injector. In the XFEL’s, only one x-ray beam is produced, and only one experiment can be performed at one time. This is not compatible with a large, multi-user community that requires many different energy ranges; and the new Canadian source will probably be based on the MBA design. Fourth-generation synchrotrons enable researchers to analyze much smaller and dilute samples of all types, and to perform in fractions of a second experiments that could take days with third-generation facilities such as CLS. Indeed, the XFEL sources have obtained spectroscopic and structural information at the femtosecond time scale on reactive intermediates in important reactions such as the manganese catalyzed production of oxygen from water in the earth’s early history, 3 billion years ago. The fourth-generation sources will also yield better spatial and spectral resolution than third-generation sources. In addition, it is likely that new spectroscopic and imaging experiments will become available. For example, the battery or catalyst chemistry described above could be monitored spectroscopically in milliseconds rather than minutes – essentially

The Positive Impact on USask and Canadian Science  141

in real time. This could well yield critical information on the chemical processes involved. The chemistry of much more dilute liquid and solid solutions of all types will be obtainable by x-ray absorption spectroscopy. The atomic structures of proteins from much smaller crystals (such as the membrane proteins mentioned previously) will be obtained more easily. For example, at the new BioMAX protein crystallography beamline at MAX IV, the x-ray beam size at the sample is 20 × 5 um2 with the remarkable photon flux of 2 × 1013 photons per second, for a 500mA beam current.54 In a different ­operating mode, MAX IV has also achieved stable 5 × 5 um2 beams. These sizes should be compared with the beam size of 130 × 25 um2 at the third-generation CLS with a photon flux of 5 × 1012 photons per second. Smaller beam sizes can only be obtained with considerably lower intensities at CLS. Many proteins change their shape and move when involved in reactions in the body. After stimulation of protein motion by a laser or electric field, the actual atomic motion of proteins can be studied on the 100 femtosecond time scale at an XFEL source or an MBA ring like MAX IV that employs a high-energy LINAC injector. The XFEL source is superior for these studies, but protein ­motion has already been demonstrated at the FemtoMAX beamline at MAX IV. The spatial resolution for imaging could be dramatically improved. Many imaging techniques, such as diffraction-enhanced imaging and the STXM previously described, can be enhanced in the fourth-generation synchrotrons, leading to dramatic improvement in resolution by up to two orders of magnitude for soft x-ray and hard x-ray imaging. One relatively new CDI technique, ptychography, can be greatly enhanced on fourth-generation synchrotrons in both the soft and hard x-ray regions. Ptychography has already achieved resolutions below 10 nanometres in the soft x-ray region and even smaller in the hard x-ray region. Moreover, ptychography provides a three-dimensional chemical image (different chemical forms of all elements) of a complex three-dimensional object, and this is sometimes termed four-dimensional imaging.55,56 Soft x-ray spectro-ptychography has already been used to characterize complex inorganic and biomaterials of all types, including bacteria, cement, nanocatalysts, working batteries, fuel cell membranes, and aerogels.56 To remain competitive, CLS and the Canadian synchrotron community more broadly must plan for a fourth-generation Canadian synchrotron for operation in the next decade. With the strong encouragement from CLS director Rob Lamb, Les Dallin, the new machine director, Mark Boland, and their co-workers at CLS have recently designed fourth-generation rings with circumferences of around 500 metres and estimated costs of many hundreds of millions of dollars. The performances that should result from these designs will exceed those of the MAX IV.57 There will have to be another concerted effort by the synchrotron community, similar to that undertaken by CISR and potential

142  The Canadian Light Source

host sites, which were crucial to getting CLS funded. Indeed, CISR (or an equivalent organization) should probably be reconstituted to mount the campaign. Initial funding of around $20 million will be required to complete the designs of the ring and some beamlines. Several workshops will have to be organized to plan the properties of the ring, the research areas, and the beamline properties. All major Canadian universities will need to be visited again to consolidate their support. Detailed letters from each university and some industries will be needed to outline the present researchers and research at CLS from each organization, as well as details of all the proposed research. The above campaign has to be started soon if a new ring is to be funded in Canada in the next ten years. It is quite possible that a competition between different potential hosts for the new facility will again occur. Perhaps this is necessary to get the community motivated. The CFI seems to be the probable funder now, and hopefully it will fund nearly the entirety of the proposal so that fundraising does not dominate the time of the director and administration, as it has for CLS.

Appendix 1

Synchrotron Facilities and Synchrotron Science: A Brief Overview

It seems useful to provide a brief description of a synchrotron facility, along with a discussion of some of the concepts and terms used in this book. References 58 and especially 59 should be consulted for more detail. Figure A1 shows a schematic of the second-generation synchrotron Aladdin outside Madison, Wisconsin. This synchrotron became operational in 1981; and three Canadian beamlines were built (called the Canadian Synchrotron Radiation Facility [CSRF], as discussed in chapter 3). The Aladdin synchrotron is a semi-circular “ring” (common to all synchrotrons) having a circumference of 89 metres, with the narrow lines shown in figure A1 indicating a metal pipe (under high vacuum) through which a beam of electrons circulate in an anti-clockwise direction (see bottom left). The beam circulates in a clockwise direction in many synchrotrons; and most current synchrotrons are much larger, with circumferences up to a few thousand metres (for example, the synchrotron at CLS has a circumference of 170 metres; see reference 34 for more details). The major components of a synchrotron facility are shown in figure A1: an electron injector, bending magnets, and beamlines coming off tangentially to the ring (see the two straight lines on the left) to utilize the synchrotron light generated by bending the electrons in the magnetic fields; a radiofrequency (RF) cavity to replenish the electron energy lost to synchrotron radiation; and an undulator (bottom of ring), which produces very brilliant synchrotron light by “wiggling” the electrons in “multipole” magnets. Many other magnets, mainly to focus the electron beam, are not shown. In the injector, the electrons are generated initially by heating a metal (such as W) to a very high temperature. The electrons are effectively “boiled off ” from the W. These emitted electrons are accelerated to very high velocities to yield high-energy electrons (25–250 MeV) in a high-vacuum tube. The injectors in different facilities are different types of accelerators: for example, a 40 MeV microtron for Aladdin and a 250 MeV linear accelerator (the SAL original accelerator) for CLS. An electron beam of approximately 200 milliamps (mA)

144  Appendix 1 Figure A1.  Schematic of the Aladdin synchrotron at the University of Wisconsin

is injected into the synchrotron (top right in figure A1). The energy of these electrons is increased to much higher energies (1 GeV [1,000 MeV] at Aladdin or 2.9 GeV at CLS), either by accelerating them further in the synchrotron itself (as at Aladdin), or in a “booster” ring (as at CLS). At 2.9 GeV, the electrons are travelling in “bunches” at more than 99.9999 per cent of the speed of light. At all times, the beam of electrons is very small – much less than a mm2 in cross-section. The fine electron beam is bent by a number of bending magnets (three on the top right corner, and twelve in total at Aladdin) to keep the electron beam in the ring. Other magnets (not shown) keep the electron beam focused so that the electron beam cross-section is very small all around the ring. The 200 to 250 mA beam decays over many hours to 100 to 150 mA due mainly to collisions with air molecules in the vacuum system. The electrons are reinjected approximately every ten hours to provide an electron beam current close to 200 mA for twenty-four-hour operation.

Synchrotron Facilities and Synchrotron Science  145 Figure A2.  Schematic of a hard x-ray beamline

Of greatest importance, when the electron beam is bent by all of the twelve bending magnets at Aladdin (or twenty-four at CLS), synchrotron light or synchrotron radiation “falls out” from all of the twelve bending magnets. This light is emitted as a very small beam in a laser-like manner tangential to the direction of motion at the middle of every bending magnet (see the two straight lines on the left side of figure A1 emitted by one of the bending magnets at Aladdin). Typically, one or two beams of light are selected from each bending magnet (see the two beams from the magnet on the left) and these beams enter a beamline, as shown in figure A2. A typical facility has twenty or more beams of light, and twenty or more beamlines to capture and use the light in experiments (see below). This emission of synchrotron light decreases the energy of the electron beam very slightly, and the microwave RF cavity (top centre) “zaps” the electrons in each revolution around the ring to keep the electron energy at a very constant value (1 GeV for Aladdin and 2.9 GeV for CLS). It should be remembered that the electrons have speeds very close to the speed of light (3 × 108 metres per second). In addition to bending magnets, other “multipole” magnets (so-called insertion devices, which are divided into undulators or wigglers; see bottom of figure A1) are now common in the straight parts of

146  Appendix 1

the ring – the so-called straight sections. These magnets generate much higher intensities of light than the bending magnets, and are the most coveted beams in any synchrotron facility. These magnets only became common in the 1990s, and so Aladdin only had two such magnets, whereas CLS has the possibility of at nine multipole magnets in the twelve straight sections. Synchrotrons like EROS at USask were initially constructed for nuclear physics experiments using the circulating electron beam, and any scientist that wanted to use the synchrotron radiation were “parasitic” users. These facilities were optimized for the electron beams and not the photon beams, and these synchrotrons are usually called “first-generation” synchrotrons. Tantalus became one of the first “second-generation” synchrotrons totally dedicated to synchrotron radiation (SR) (often called a storage ring, because the electron beam was stored for many hours before reinjection of the electrons) that was optimized for the photon beams, and used only the photon beams for experiments. Aladdin was a higher-energy and higher-performance second-generation synchrotron compared to Tantalus. When special magnets, such a wigglers and undulators in the straight sections (see figure A1) began to dramatically enhance the intensity and quality of the photon beams, synchrotrons such as CLS began to be constructed in the early to mid-1990s in many countries. These are referred to as “third-generation” facilities. Fourth-generation synchrotrons are now being constructed in many countries with much smaller and intense laser-like photon beams. CLS will probably be converted to a fourth-generation ring in the next ten years (as discussed at the end of chapter 8). Every synchrotron around the world is unique; and as usual, the budget largely determines the quality (i.e., intensity and brilliance) of the available synchrotron beams. In order to make the CLS synchrotron fit into a relatively small building (about 100 metres by 100 metres – much smaller than other similar facilities), make the facility affordable, have a large number of possible beamlines, and be competitive for intensity and brilliance of the light beams, novel magnets were designed at SAL that would both bend and focus the electron beams. As a result, the CLS facility is probably the most cost-effective synchrotron in the world – a great credit to Dennis Skopik, Les Dallin, and their team. For example, the cost of the CLS facility was less than one-half of other competitive facilities, such as the Diamond synchrotron in England. But why are these beams of light so important for a myriad of research uses? Each beam is incredibly intense and contains “light” with frequencies covering more than one-half of the electromagnetic spectrum – from the low-energy far-infrared through the infrared, visible, ultraviolet, soft x-ray, and hard x-ray regions. Beamlines are constructed to optimize performance in the different energy regions (except for the visible region), and each synchrotron facility has one or more beamlines in each region. The hard x-ray region is perhaps the most important in most new facilities such as CLS (the ultraviolet and soft

Synchrotron Facilities and Synchrotron Science  147

x-ray regions were most important at Aladdin, and the infrared and soft x-ray regions were most important for the initial seven beamlines at CLS), and a schematic of a hard x-ray beamline is shown in figure A2. These beamlines all use high-vacuum conditions in a metal pipe to transport the beam. At the top right of figure A2, the synchrotron beam is emitted when the electrons are bent. This beam contains the broad spectrum of energies noted above. In nearly every beamline, a monochromator is employed to give a single energy and to vary the energy over a predetermined energy spread, for example from 2 keV to 10 keV. For x-ray beamlines, two crystals are used to monochromatize (create one energy) the beam via diffraction by two crystals and to vary its energy. Depending on the nature of the two crystals, different energy ranges can be obtained (e.g., between 2 keV and 10 keV with one type of crystal in one beamline, and between 4 keV and 50 keV with another crystal in another beamline). The beam is then focused with one or two mirrors (shown in figure A2) before it hits a sample in an experimental chamber. Such chambers are often very complex, with a number of detectors, depending on the type of experiment. Both the monochromator and the experimental detectors are controlled by a computer that can often store gigabytes of data. Because of the very high intensities and readily variable energies, scientists from nearly all scientific areas (e.g., physics, chemistry, biology, geology, medicine, and engineering), need SR to perform critical experiments in their field. It would require another large book to look at all the uses of SR (see references 1–3, 34, and especially 59), but a few examples in chapter 8 of this book will hopefully give the reader a good idea of the scope of the research possible using SR. For further information, please look at the CLS website58 and especially a recent book.59 It is perhaps interesting that nearly all the technology described above in a synchrotron facility uses the pure scientific discoveries of the last 120 years. I took a “physics of the atom” course in 1961 at the University of Manitoba. This course fascinated me, and led directly to my career in spectroscopy. The electron was discovered in “cathode rays” in 1897 by J.J. Thomson at Cambridge; and I was always inspired by this great discovery when I was doing my PhD studies at the University of Cambridge, partly because I was resident in the Sir George Thomson building, so named after J.J.’s son. (And as a matter of fact, both father and son won the Nobel Prize in Physics – J.J. in 1906 and George in 1937.) J.J. Thomson soon measured many properties of the electron; for example, the charge to mass ratio q/m, and the deflection of the electron in electric and magnetic fields. These discoveries, of course, are critical for the acceleration and bending of electrons. For high-energy electrons (even those electrons in a cathode ray tube), the mass of an electron increases at very high energies and velocities due to relativistic effects explained in Einstein’s early work in the 1900s. To compute the path of a high-energy electron in the synchrotron,

148  Appendix 1

the relativistic change in mass is readily computed from the original, simple Einstein equations. X-rays are even more important than electrons when talking about synchrotron facilities, and they were discovered earlier than the electron by Roentgen in 1895. They were soon characterized by diffraction off both ruled gratings and crystals. In 1912, Bragg discovered the scattering of x-rays by atoms in a crystal, and this led quickly to the huge field of x-ray crystallography, which enabled the determination of the atomic structure of any crystal – and of course to the simple diffraction equations necessary for designing the double crystal monochromator (shown in figure A2). Only in the late 1950s was this technique used to determine the atomic structure of very large molecules such as proteins – a major use of synchrotron facilities today. Without computers at the time, such structures were very difficult to solve. But SR was not discovered until 1947, at an accelerator at the General Electric Lab in Schenectady, New York. It is interesting to note that a Canadian, John Blewett (who graduated from both the University of Toronto and Princeton) was one of the discoverers. This novel radiation source was not used until the late 1960s, when it was utilized in a few labs at Stanford, Wisconsin, Hamburg, and Tokyo. By the time I heard about SR in the early 1970s, it was becoming very clear that this radiation would be extremely important scientifically.

Appendix 2

Canadian Institute for Synchrotron Radiation: Announcement of CFI Funding, 1999 g . m . bancroft , president , june 3, 1999

Finally! Amazing! For our tenth annual CISR meeting, we now have funding for the Canadian Light Source (CLS). On Wednesday, March 31, 1999, The CFI announced a $56M contribution towards building the $173.5M CLS synchrotron project at the University of Saskatchewan in Saskatoon. As the press announcement from Saskatoon said: “CLS represents an unprecedented level of collaboration among governments, Universities and industry in Canada.” Eighteen Universities, in addition to the University of ­Saskatchewan, have endorsed the CLS project on behalf of over 300 users of synchrotron light in Canada. Before elaborating on this quite incredible level of cooperation among a very large number of very talented people, I want to summarize the major events of the last year that led to the CFI announcement. On May 28, 1998, the ­final CLS proposal was submitted to the first CFI competition. It received very high ratings, and the proposal was advanced to the second round of competition in October 1998. On Feb. 22 and 23, 1999 in Saskatoon, an international committee with representatives from foreign synchrotrons such as APS, Soleil, and ESRF, reviewed the economic viability of the project. The previous 1996 NSERC committee had strongly recommended the project scientifically, so ­little time was spent on the science. The review went well, but the CFI committee obviously wanted firmer operating commitments from NRC and NSERC. Firmer commitments were given to the CFI following the meeting. The CFI Multidisciplinary Assessment Committee (MAC) met to review the proposals on March 8, 9, and 10, 1999. Both the Site Review Committee and the MAC recommended funding for CLS; and on March 31, the CFI announced FULL funding as requested. Obviously, a tremendous amount of work went into the submissions and reviews in the last year. We all have to give Dennis Skopik an enormous vote of thanks for guiding the overall project in a VERY complex environment. Without his administrative skills and experience, this project would not have

150  Appendix 2

happened. CISR members were very important for generating many of the ­University letters of support. Your executive (Adam Hitchcock, Ron Cavell, T.K. Sham, and I) worked very hard with Dennis to get the application together, and Dennis constantly consulted us on policy and monetary issues. Dennis Skopik, Dennis Johnson, and I made many visits to Universities and industries to garner support. All of us will be very sorry to see Dennis Skopik leave the project for the CEBAF lab in Virginia. But this project required the tremendous co-operation and drive of many others from the University of Saskatchewan, the city of Saskatoon, the ­Saskatchewan government, and the federal government. George Ivany, president of the University of Saskatchewan, did a magnificent job in getting the university and province on side. There was not another university president in Canada who had such vision, enthusiasm, and drive. Michael Corcoran (VP Research), Tony Whitworth (VP Finance), Dennis Johnson (former VP Research) and the whole SAL staff were always working hard on this project. I would like to single out Emil Hallin, who has done a yeomen’s job all by himself on the beamline part of the project. Barry Hawkins, as Project Manager from UMA, has been incredibly important recently in ­organizing the project. The Collaborative Committee (Bernard Michel, Chair; Hal Wyatt, Chair of the University Board of Governors; Fraser Nicholson and Larry Spanier, Saskatchewan Ministry of Economic and Co-operative Development; Onno Kremers and Doug Maley, Western Diversification) were very important in promoting the project. Clive Willis from NRC (supported strongly by ­Arthur Carty, ­President of NRC) worked effectively with the CISR executive in ­lobbying, helping to organize the five successful industrial workshops from ­ December 1997 to April 1998, and effecting policy. Two university lawyers from Mckercher Mckercher and Whitmore, Doug Richardson and Dick B ­ atten, were very important for political lobbying provincially and federally, and in crafting several legal documents such as the government structure of CLS. Roy Romanow (Premier of Saskatchewan), Henry Dayday (Mayor of ­Saskatoon) and Janice MacKinnon (Minister of Economic and Co-operative Development) contributed government resources and lobbied hard for the project. Both Roy and Janice were kind enough to send me a congratulatory letter after the CFI announcement. Larry Spannier has been important in guiding and funding the project effectively in the last year. Peter Wyant from Crown investments did an excellent job on all the budgets and budget details. Federally, we must thank Ralph Goodale, Saskatchewan MP and Minister of Natural Resources, for a great deal of effort raising federal funds, especially when in-kind laboratory matching funds were not allowed by the CFI. The Western Diversification staff (Doug Maley and Onno Kremers) have already

Announcement of CFI Funding, 1999  151

been mentioned. We also have to thank Peter Morand, the former President of NSERC, Tom Brzustowski, current President of NSERC, Arthur Carty, ­President of NRC, and Henry Friesen, President of MRC, for their generous support, financially, scientifically, and politically. Norman Sherman and Walter ­Davidson from NRC have also been very helpful. There are many scientists – Canadian, US, and European – who have sat on NSERC committees and received proposals for review, who have supported our effort enthusiastically. For example, Bob McAlpine from NSERC, and John Strom-Olsen and Alex McAuley (and their committees), spent countless hours on behalf of the CLS project. I should also acknowledge Bruce Bigham, recently retired from AECL, who initiated CISR in 1990. Also the staff at CSRF (Kim Tan, Brian Yates, Emil ­Hallin, Greg Retzlaff, X.H. Feng, and B.X. Yang) built the three beamlines at CSRF in Madison and increased the interest greatly in SR applications in the last fifteen years. (As an aside, it is really important to thank Paul Martin, Federal Minister of Finance, for setting up the CFI out of the Finance Ministry – a unique event. Without the CFI, CLS, and a large number of Canadian research projects and professorships would not have been funded. Canadian universities owe Paul Martin and the CFI an enormous vote of thanks.) Now the real work begins and we must now deliver! The ring is nearly ­finalized (a special vote of thanks to Les Dallin), but there is much to do on the beamlines, not the least of which is to raise $19M more of the $48M for beamlines. Already, Tom Ellis has organized a very successful IR workshop in Montreal. Three IR beamlines are being contemplated, with three strong groups to help organize and design the beamlines and raise external funds. A document has been written by Adam Hitchcock and me for “Procedures to Request, Review, and Approve CLS Beamline Projects.” It is hoped that several other beamline projects will begin organization at this meeting. Most of the beamline organization should be in place this autumn, and the beamlines must be designed by next summer if they are to be operational by the end of 2003. It is time for me now to step down as President of CISR. It has been a great adventure with many talented people. I hope to continue the adventure in ­Saskatoon (my birthplace) as director (pro tem) in September.

References

1) Klein, G. “Synchrotron: Canadian Light Source 70 years in the making.” Saskatoon Star Phoenix, Oct. 20, 2004. 2) Bancroft, G.M. “The Canadian Light Source: Progress and prospects.” Canadian Chemical News 53 (2001): 18–21. 3) Bancroft, G.M. “The Canadian Light Source: History and scientific prospects.” Canadian Journal of Chemistry 82, no. 6 (2004): 1028–42. https://doi.org/10.1139 /v04-027 4) Bancroft, G.M. “The Canadian Synchrotron Radiation Facility (CSRF) in Madison: Twenty-five years of soft x-ray research.” Canadian Journal of Chemistry 85, no. 10 (2007) 637–44. https://doi.org/10.1139/v07-047 5) Stoicheff, B. Gerhard Herzberg: An illustrious life in science. Ottawa: NRC Press, 2002. 6) Herzberg, G. Atomic spectra and atomic structure. New York: Dover Publications, 1944. 7) Houston, C. Stuart. Steps on the road to Medicare: Why Saskatchewan led the way. Montreal: McGill-Queen’s University Press, 2002. 8) Harrington, E.L., R.N.H. Haslam, H.E. Johns, and L. Katz. “The betatron building and installation at the University of Saskatchewan.” Science 110 (1949): 283–5. https://doi.org/10.1126/science.110.2855.283 9) Johns, H.E., L. Katz, R.A. Douglas, and R.N.H. Haslam. “Gamma-neutron cross-sections.” Physical Review 80 (1950): 1062. https://doi.org/10.1103/ PhysRev.80.1062 10) Hayden, M. Seeking a balance: University of Saskatchewan, 1907–1982. Vancouver: University of British Columbia Press, 1983. 11) Katz, L., G.A. Beer, D.E. McArthur, H.S. Caplan. “The electron-scattering facility at the Saskatchewan Accelerator Laboratory.” Canadian Journal of Physics 45 (1967): 3721–36. https://doi.org/10.1139/p67-311 12) MacKinnon, P. University leadership and public policy in the twenty-first century. Toronto: University of Toronto Press, 2014. 13) Norum, B.E., J.C. Bergstrom, and H.S. Caplan. “Electroexitation of the giant resonance of 17O.” Nuclear Physics A 289 (1977): 275–91. https://doi .org/10.1016/0375-9474(77)90033-1

154 References 14) Bergstrom, J.C., H.S. Caplan, R.V. Servranckx, and B.E. Norum. “SORE – A pulse stretcher for the Saskatchewan 300 MeV linac.” IEEE Transactions on Nuclear Science 30, no. 4 (1983): 3226–8. https://doi.org/10.1109/TNS.1983.4336623 15) Vogt, J.M., R.E. Pywell, D.M. Skopik, E.L. Hallin, J.C. Bergstrom, H.S. Caplan, K.I. Blomqvist, and W. Del Bianco. “The photon tagging facility at the Saskatchewan Accelerator Laboratory.” Nuclear Instruments and Methods in Physics Research Section A 324 (1993): 198–208. https://doi.org/10.1016/0168-9002(93)90977-P 16) Vogt, J.M., J.C. Bergstrom, R. Igarashi, and K.J. Keeter. “Igloo: A neutral pion spectrometer for low energy photoproduction studies.” Nuclear Instruments and Methods in Physics Research Section A 366 (1995): 100–14. https://doi .org/10.1016/0168-9002(95)00557-9 17) Dallin, L.O. “Operating results of the electron ring of Saskatchewan (EROS).” Proceedings of the 1989 IEEE Particle Accelerator Conference (1989): 21–6. https:// doi.org/10.1109/PAC.1989.73026 18) Bancroft, G.M. Mössbauer spectroscopy: An introduction for inorganic chemists and geochemists. London: McGraw-Hill, 1973. 19) Turner, D.W., C. Baker, A.D. Baker, and C.R. Brundle. Molecular photoelectron spectroscopy. New York: Wiley, 1970. 20) Siegbahn, K., C. Nordling, G. Johansson, J. Hedman, P.F. Heden, K. Hamrin, U. Gelius, T. Bergmark, L.O. Werme, R. Manne, and Y. Baer. ESCA applied to free molecules. Amsterdam: North Holland Press, 1971. 21) Bancroft, G.M., and P.W.M. Jacobs. “Some reasons why Canadian scientists should have their own synchrotron.” Science Forum (Dec. 1975): 21–3. 22) Hitchcock, A.P., and G.M. Bancroft. “A unique, rapidly developing tool for chemical research.” Chemistry in Canada 43 (1991): 16–20. 23) Crozier, E.D., F.W. Lytle, D.E. Sayers, and E.A. Stern. “Structural determinations of liquid semiconductors using extended x-ray absorption fine structure.” Canadian Journal of Chemistry 55 (1977): 1968–74. https://doi.org/10.1139 /v77-274 24) Lynch, D.W., W. Plummer, F. Himpsel, T.C. Chiang, G. Margaritando, and G. Lapeyre. “Tantalus, the first dedicated synchrotron source.” Synchrotron Radiation News 28, no. 4 (2015): 20–3. https://doi.org/10.1080/08940886.2015.1059232 25) Bancroft, G.M., T.K. Sham, D.E. Eastman, and W. Gudat. “Photoelectron spectra of solid inorganic and organometallic compounds using synchrotron radiation: Valence band spectra and ligand field splitting: Broadening of core d levels.” Journal of the American Chemical Society 99 (1977): 1752–62. https://doi. org/10.1021/ja00448a012 26) Bancroft, G.M., P.A. Malmquist, S. Svensson, E. Basilier, U. Gelius, and K. Siegbahn. “Gas phase ESCA studies of valence and core levels in XeF2 and XeF4.” Inorganic Chemistry 17 (1978): 1595–1600. https://doi.org/10.1021/ic50184a040 27) Tan, K.H., G.M. Bancroft, L.L. Coatsworth, and B.W. Yates. “Mark IV ‘Grasshopper’ grazing incidence monochromator for the Canadian Synchrotron

References 155

28)

29)

30)

31) 32)

33)

34)

35)

36)

37)

38)

Radiation Facility (CSRF).” Canadian Journal of Physics 60 (1982): 131–7. https:// doi.org/10.1139/p82-018 Yates, B.W., K.H. Tan, G.M. Bancroft, L.L. Coatsworth, and J.S. Tse. “Photoelectron study of the valence level cross sections of CF4 and SiF4 from 21 eV to 100 eV using synchrotron radiation.” Journal of Chemical Physics 83 (1985): 4906–13. https://doi.org/10.1063/1.449749 Leung, W.M., D.M. Shinozaki, and J.W. McGowan. “Soft x-ray contact microscopy of polyethylene microstructures.” Journal of Materials Science 20 (1985): 46–52. https://doi.org/10.1007/BF00555897 Aksela, H., S. Aksela, G.M. Bancroft, K.H. Tan, and H. Pulkkinen. “Study of the N4,5OO resonance auger spectra of Xe using selective excitation by synchrotron radiation.” Physical Review A 33 (1986): 3867–75. https://doi.org/10.1103/PhysRevA.33.3867 Bancroft, G.M. “New developments in far UV, soft x-ray research at the Canadian Synchrotron Radiation Facility.” Canadian Chemical News 44 (1992): 15–22. Yang, B.X., F. H. Middleton, B.G. Olsson, G.M. Bancroft, J.M. Chen, T.K. Sham, K.H. Tan, and D. Wallace. “The design and performance of a soft x-ray double crystal monochromator beamline at Aladdin.” Nuclear Instruments and Methods in Physics Research Section A 316 (1992): 422–36. https://doi.org/10.1016/0168 -9002(92)90930-3 Yates, B.W., Y.F. Hu, K.H. Tan, G. Retzlaff, R.G. Cavell, T.K. Sham, and G.M. Bancroft. “First results from the Canadian SGM beamline at SRC.” Journal of Synchrotron Radiation 7 (2000): 296–300. https://doi.org/10.1107/S0909049500007214 Cutler, J., D. Chapman, and R. Lamb. “Brightest light in Canada: The Canadian Light Source.” Synchrotron Radiation News 31 (2018): 26–31. https://doi.org/10.10 80/08940886.2018.1409557 Li, X., M. Banis, A. Lushington, X. Yang, Q. Sun, Y. Zhao, C. Liu, Q. Li, B. Wang, W. Xiao, C. Wang, M. Li, J. Liang, R. Li, Y. Hu, L. Goncharova, H. Zhang, T.K. Sham, and X, Sun. “A high-energy sulfur cathode in carbonate electrolyte by eliminating polysulphides via solid phase lithium-sulfur transformation.” Nature Communications 9 (2018): 4509–19. https://doi.org/10.1038/s41467-018-06877-9 Ghirenghelli, G., M. Le Tacon, M. Minola, S. Blanco-Canosa, C. Mazzoli, N.B. Brookes, G.M. De Luca, A. Frano, D.G. Hawthorne, F. He, T. Loew, M. Moretti Sala, D.C. Peets, M. Salluro, E. Schierle, R. Sutarto, G.A. Sawatzky, E. Weschke, B. Keimer, and L. Braicovich. “Long-range incommensurate charge fluctuations in (Y,Nd) 6+x.” Science 337 (2012): 821–5. https://doi.org/10.1126/science.1223532 Comin, R., R. Sutarto, E.H. da Silva Neto, L. Chauviere, R. Liang, W.N. Hardy, D.A. Bonn, F. He, G.A. Sawatzky, and A. Damascelli. “Broken translational and rotational symmetry via charge stripe order in underdoped YBa2Cu3O6+y.” Science 347 (2015): 1335–9. https://doi.org/10.1126/science.1258399 Liang, Y., Y. Li, H. Wang, J. Zhou, J. Wang, T. Regier, and H. Dai. “Co3O4 nanocrystals on graphene as a synergistic catalyst for oxygen reduction reaction.” Nature Materials 10 (2011): 780–6. https://doi.org/10.1038/nmat3087

156 References 39) Bo, Z., X. Zheng, O. Vozvyy, R. Conin, M. Bajdich, M. Garcia-Melchor, L. Han, J. Lu, M. Liu, L. Zheng, F. Pelayo Garcia de Arquer, C.T. Dinh, F. Fan, M. Yuan, E. Yassetepe, N. Chen, T. Regier, P. Liu, Y. Li, P. de Luna, A. Janmohamed, H.L. Xin, H.Yang, A. Vojvodic, and E.H. Sargent. “Homogeneously dispersed multimetal oxygen-evolving catalysts.” Science 352 (2016): 333–7. https://doi.org/10.1126 /science.aaf1525 40) Grochulski, P., M.Fodje, S. Labiuk, T.W. Wysokinski, G. Belov, M. Korbas, and S.M. Rosendahl. “Review of Canadian Light Source facilities for biological applications.” Nuclear Instruments and Methods in Physics Research 411 (2016): 17–21. https://doi.org/10.1016/j.nimb.2017.01.065 41) Zamyatkin, D.F., F. Parra, J.M. Martin Alonso, D.A. Harki, B.R. Peterson, P. Grochulski, and K.K.S. Ng. “Structural insights into mechanisms of catalysis and inhibition in Norwalk virus polymerase.” Journal of Biological Chemistry 283 (2008): 7705–12. https://doi.org/10.1074/jbc.M709563200 42) Grandin, M., M. Meier, J.G. Delcros, D. Nikodemus, R. Reutin, T.R. Patel, D. Goldschneider, G. Orris, N. Krahn, A. Boussouar, R. Abes, Y. Dean, D. Neves, A. Bernet, S. Depil, F. Schneiders, K. Poole, R. Dante, M. Koch, P. Meglen, and J. Stetefeld. “Structural decoding of the Netrin-1/UNC5 interaction and its therapeutical implications in cancers.” Cancer Cell 29 (2016): 173–85. https://doi .org/10.1016/j.ccell.2016.01.001 43) Telling, N.D., J. Everett, J.F. Collingwood, J. Dobson, G. van der Laan, J.J. Gallagher, J. Wang, and A.P. Hitchcock. “Iron biochemistry is correlated with amyloid plaque morphology in an established mouse model of Alzheimer’s disease.” Cell Chemical Biology 24 (2017): 1205–15. https://doi.org/10.1016 /j.chembiol.2017.07.014 44) Luan, X., V.A. Campanucci, M. Nair, O. Yilmaz, G. Belov, T.E. Machen, L.D. Chapman, and J.P. Ianowski. “Pseudomas Aeruginosa triggers CFTR-mediated airway surface liquid secretion in swine trachea.” Proceedings of the National Academy of Sciences 111 (2014): 12930–5. https://doi.org/10.1073 /pnas.1406414111 45) Luan, X., G. Belov, J.S. Tam, S. Jagadeeshan, N. Hassan, P. Giono, N. Grishchenko, Y. Huang, J.L. Carmalt, T. Duke, T. Jones, B. Monson, M. Burmester, T. Simovich, O. Yilmaz, V.A. Campanucci, T.E. Machen, L.D. Chapman, and J.P. Ianowski. “Cystic fibrosis swine fail to secrete airway surface liquid in response to inhalation of pathogens.” Nature Communications 786 (2017). https://doi.org/10.1038 /s41467-017-00835-7 46) Barber, A., J. Brandes, A. Leri, K. Lalonde, K. Balind, S. Wirick, J. Wang, and Y. Gelian. “Preservation of organic matter in marine sediments by inner-sphere interactions with reactive iron.” Scientific Reports 7, no. 1 (2017): 366–76. https:// doi.org/10.1038/s41598-017-00494-0 47) Chen, N., D.T. Jiang, J.N. Cutler, T. Kotzer, Y.F. Jia, G.P. Demopoulos, and J.W. Rowson. “Structural characterization of poorly-crystalline scorodite, iron

References 157

48)

49)

50)

51) 52) 53)

54) 55)

56)

57)

58) 59)

(III)-arsenate co-precipitates and uranium mill neutralized raffinate solids using X-ray absorption fine structure spectroscopy.” Geochimica et Cosmochimica Acta 73 (2009): 3260–76. https://doi.org/10.1016/j.gca.2009.02.019 Blanchard, P.E.R., L.L. Van Loon, J.W. Reid, J.N. Cutler, J. Rowson, K.A. Hughes, C.B. Brown, J.J. Mahoney, L. Xu, M. Bohan, and G.N. Demopoulos. “Investigating arsenic speciation in the JEB tailings management facility at McLean Lake, Saskatchewan using x-ray absorption spectroscopy.” Chemical Geology 466 (2017): 617–26. https://doi.org/10.1016/j.chemgeo.2017.07.014 Hestrin, R., D. Torres-Rojas, J.J. Dynes, J.M. Hook, T.Z. Regier, A.W. Gillespie, R.J. Smernik, and J. Lehmann. “Fire-derived organic matter retains ammonia through covalent bond formation.” Nature Communications 10, no. 1 (2019): 1038–46. https://doi.org/10.1038/s41467-019-08401-z Rachid, R., T. Song, M. Chu, F. Yu, S. Kumar, C. Karunakaran, and G. Peng. “Evaluating changes in cell-wall components associated with clubroot resistance using Fourrier transform infrared spectroscopy and Pt-PCR.” International Journal of Molecular Sciences 18, no. 10 (2017): 2058–72. https://doi.org/10.3390/ijms18102058 Woodhouse, H. Selling out: Academic freedom and the corporate market. Montreal: McGill-Queen’s University Press, 2009. Canadian Isotope Innovations Inc. website: http://isotopeinnovations.com Kozachuk, M.S., T.K. Sham, R.R. Martin, A.J. Nelson, I. Coulthard, and John McElhone. “Recovery of degraded-beyond-recognition 19th century Daguerreotypes with rapid high dynamic range elemental x-ray fluorescence imaging of mercury L emission.” Scientific Reports 8 (2018): 9565–72. https://doi .org/10.1038/s41598-018-27714-5 MAX IV website: http://www.maxiv.lu.se Dierolf, M., A. Menzel, P. Thibailt, P. Schneider, C.M. Kewish, R. Wepf, O. Bunk, and F. Pfeiffer. “Ptychographic x-ray computed tomography at the nanoscale.” Nature 467 (2010): 436–40. https://doi.org/10.1038/nature09419 Wu, J., X. Zhu, D. Shapiro, J.R.I. Lee, T. Van Buren, M.M. Biener, S.A. Gammon, T.T. Li, T.F. Baumann, and A.P. Hitchcok. “Four dimensional imaging of ZnOcoated alumina aerogels by scanning transmission x-ray microscopy and ptychography tomography.” Journal of Physical Chemistry 122 (2018): 25374–85. https://doi.org/10.1021/acs.jpcc.8b07363 Dallin, L., and W. Wurtz. “Towards a 4th generation storage ring at the Canadian Light Source.” Proceedings of the 12th International Conference on Synchrotron Radiation Instrumentation, SRI2015, AIOP conf. proc. 1741, 020034-1-020034-4, 2015. CLS website: http://www.lightsource.ca/inside_the_synchrotron Mobilio, S., F. Boscherini, and C. Meneghini, eds. Synchrotron radiation: Basics, methods and applications. Berlin and Heidelberg: Springer Publishing, 2014.

Index

The letter f following a page number denotes a figure; the letter n a footnote; the letter t a table. Addison, Brenda, 33f Adlington, Al, 53 Advanced Light Source (ALS), 6n, 42, 67, 98 advanced materials research, 3, 6, 101, 125–9; batteries, 125–6, 127–8, 140–1; catalysts, 129, 140–­1; superconductors, 126–7, 128–9 Advanced Photon Source (APS), 6n, 42, 56, 67, 85, 111, 131, 149 Advanced Technology Association, 67 Advisory Committee on Site Selection for the CLS Proposal, 55–6 Ahluwalia, Pardeep, 43 Aksela, Seppo and Helena, 33 Aladdin synchrotron, 21–2, 29, 32, 34– 36, 39, 40–1, 40f, 42, 57, 63, 98, 139; description, 143–7, 144f, 145f Alberta Synchrotron Institute, 91 Alcock, Reg (MP), 50–1 Aldridge, Gary, 50 Alexei, Ken, 91 American Philosophical Society, 9 Anderson, Paul, 92 Appathurai, Narayan, 38–9, 102 Argonne National Laboratory (Chicago), 42 Armour, Catherine, 84 Atkinson, Michael, 98–100

Atomic Energy Control Board of Canada, 10, 11n Atomic Energy of Canada Ltd. (AECL), 24, 29, 36, 42–3 Axworthy, Lloyd (MP), 67 Bacon, David, 44 Baker Engineering, 32, 35. See also Grasshopper monochromator Baldwin, Howard, 31 Bancroft, Catherine, 26, 99, 162 Bancroft, David, 26, 99, 162 Bancroft, Eileen, 88, 102 Bancroft, Joan, 2, 26–8, 87, 102, 124, 162 Bancroft, John, 29–31, 61 Bancroft, Mike, 33f, 37f, 40f, 84, 103f Banis, Mohammed, 127–8 Barnes, Sue (MP), 56–7 Batten, Dick, 49, 76–8, 92–3, 95, 150 Beam Inc., 75–9, 83, 85, 96 Beamline Development Inc. See Beam Inc. Beamline Planning and Access Committee (BPAC), 96–7 beamline teams (BTs), 96–7, 103 Bednorz, Georg, 128 Bergman, Katherine, 99 Bergstrom, Jack, 13–14, 16–17, 60

160 Index Bernier, Maxine: People’s Party of Canada, 123 Berutti, Franco, 48 Bigham, Bruce, 19, 42–3, 151 Blaine, Isabelle, 46 Blair, Allan, 9–10 Blewett, John, 148 Bogart, David, 87, 89–90, 108 Boland, Mark, 141 Botting, Dale, 99 Bozek, John, 35 Brain and Mind Institute, 119 Bridger, Bill, 52, 56, 95 Brion, Chris, 28, 31 Broda, Anne, 99 Brookhaven National Laboratory, 35–6, 42 Brown, Fred, 32 Brown, Gordon, 100 Brown, Jim, 133 Brzustowski, Tom, 44, 60, 63, 69, 85, 151 Burns, Roger, 23–4 Bzowski, Arthur, 52–5 Caldwell, Glen, 12, 52–5 Calvert, Lorne (prem. Sask.), 103 Canada Excellence Research Chairs (CERCs), 119 Canada Foundation for Innovation (CFI), 2, 5–6, 46, 71, 73–80, 82–5, 94, 100, 106–21; award agreement, 85–6; background, 106–10; funding support, 20, 43, 73–8, 80–3, 84t, 87–8, 101–2, 105–6, 106t, 121, 142, 149–51; Multidisciplinary Assessment Committee (MAC), 85, 149; ­USask-CFI contract, 98 Canada Research Chairs (CRCs), 81, 97, 100–1, 105–6, 108–9, 119 Canadian Association of Physicists (CAP), 107

Canadian Institute for Advanced Research, 129 Canadian Institute for Neutron Scattering, 42 Canadian Institute for Synchrotron Radiation (CISR), 4, 42–5, 47, 49, 55, 58, 60–1, 63, 65, 68–9, 71–2, 76, 78–9, 81–2, 121, 141–2, 149–51 Canadian Institutes of Health Research (CIHR), 51, 108–9, 116. See also Medical Research Council (MRC) Canadian Isotope Innovations (CII) Inc., 5, 11n, 45n, 109, 136, 138 Canadian Macromolecular Crystallography Facility, 131 Canadian Nuclear Safety Commission (CNSC), 93–4, 111 Canadian Society for Chemistry (CSC), 43, 107 Canadian Synchrotron Radiation Facility (CSRF), 4, 21–41, 33f, 37f, 39f, 40f, 42–72, 122, 143; background, 21–2; funding, 21, 29–31, 33–5, 36–8, 40, 50, 55, 57, 61; in Madison, Wisconsin, 28–35, 42, 77, 79, 89–90, 102, 122–4, 133, 151 Caplan, Henry, 16, 19 Carnegie Foundation, 9 Carr, Jack, 102 Carter, Lavina, 102 Carty, Arthur, 31, 51, 67, 69, 77–8, 85, 95, 101–3, 150–1 Cavell, Ron, 24, 35, 37f, 40, 43, 60, 84, 90, 95, 97, 150 Centre for Chemical Physics (CCP), 4, 23–5, 29–31, 33, 45, 124, 161 Chad, Karen, 119 Chakma, Amit, 120 Chalk River Nuclear Laboratories, 19, 29, 94, 137 Chapman, Dean, 100, 132 Charette, Carmen, 46

Index 161 Chen, Jin-Ming, 35–6, 122 Chen, Ning, 103, 129 CHESS synchrotron (Cornell), 135 Childers, C.E., 70 Childs, Ron, 56 China Automotive Battery Research Institute Co. Ltd., 128 Chrétien, Jean (PM), 5, 44, 102, 103f, 106 Clark, Howard, 24, 31, 52–3, 57 Clark, Michael, 23 Clark, Scott, 106 CLS (Canadian Light Source): and CFI, 121; and convergence, 126–36; economic impact, 115–17, 116f, 136–7; educational programs, 138–9; facility description, 112, 138; funding, 82–5, 84f, 87–92, 102, 105–6, 106t, 121, 149–51; impact on USask, 115–21; and internationalization, 125–6; origins, 64–8; Strategy and Action Plan, 65–8, 75; Team Canada, 65–71, 75; usage, 111–12, 136–7; users, 112, 113tt, 114t, 115, 136–7 CLS Group. See USask Steering Committee CLS Inc., 75, 77–9, 83, 85, 93–4 CLS organizational structure, 111; Board, 78, 89, 93, 95, 97, 103, 112, 115, 137–8; Facility Advisory Committee (FAC), 96, 100; Industrial Committee, 96; Machine Advisory Committee (MAC), 96; President’s Committee, 91, 94–7; Review Oversight Committee (ROC), 96; Science Advisory Committee (SAC), 96; Users Advisory Committee (UAC), 96–7 CLS Strategy and Action Plan, 65–8, 75 Coatsworth, Leighton, 29 Cockcroft, Sir John, 12, 13f

Collaborative Committee, 74–80, 82, 150 Connell, George, 29–31, 56–7, 70 Cooper, David, 100 Corcoran, Michael, 76–7, 81, 84, 95–6, 150 Coulthard, Ian, 38–9, 40f, 102–3 Coyne, Andrew, 123 Craddock, Michael, 43 Craig, Wayne, 99 Crozier, Daryl, 6n, 25, 31, 38, 39f, 43, 60, 111 Cunningham, Diane (MPP), 54, 56–7 Currie, Gavin, 124 Cutler, Jeff, 35, 38–9, 40f, 53–5, 102, 132–3, 136 Cygler, Miroslaw, 100 Dai, H., 129 Dalai, Ajay, 100 Dallin, Les, 17, 17f, 43, 50, 57, 60, 122, 141, 146, 151 Davenport, Paul, 45, 51–6, 58, 70, 86 Davidson, Walter, 12n, 31, 37f, 38, 84, 95, 137, 151 Davies, Paul, 124 Dayday, Henry, 51, 59, 150 Delbaere, Louis, 44, 60, 98, 100 Diamond synchrotron (UK), 146 Dingwell, David (MP), 67 Domascelli, Andrei, 129 Dombowsky, David, 63–4, 66–8 double crystal monochromator (DCM), 35–6, 38, 148 Douglas, Alec, 24 Douglas, Tommy (prem. Sask), 11 DRI Canada report, 69 Duhamel, Ron (MP), 80 Eastman, Dean, 25–6 Eggelton, Art (MP), 67 Electron Ring of Saskatchewan (EROS), 14–17, 58–60, 146; funding, 15

162 Index Ellis, Tom, 57, 98, 139, 151 environment and agriculture research, 6, 11, 101, 113t, 126, 132–6; ammonia emissions, 135; clubroot resistance, 135–6; marine sediments, 133–4; toxic arsenic, 134 Erickson, Gloria, 88 European Synchrotron Radiation Facility (ESRF), 84–5, 149 Fairbairn, Joyce, 67 Feasby, Tom, 31 Feng, X.H., 151 Filmon, Gary (prem. Manitoba), 92; Janice (spouse), 92 Fleet, Mike, 57 Flumian, Maryantonett, 75 Foster, Gordon, 124 Franklin, Colin, 84 Fraser, Marie, 60 Friedman, Thomas, 122 Friesen, Henry, 151 Froelich. Hans, 16 Fuller, Marina, 36 Gagnon, Denis, 84 Gélinas, Yves, 134 George, Graham, 100 Gerrard, Jon (MP), 50–1, 67, 70, 72–4 GLABAT Solid State Battery Inc., 128 Goodale, Ralph (MP), 44, 46–7, 50–1, 67, 74, 80, 103, 150 Grasshopper monochromator, 32–5, 33f, 38–40, 40f. See also Baker Engineering Green, Mike, 52–3, 57 Griffiths, Keith, 29 Groat, Lee, 60 Grochulski, Pawel, 103, 131 Grosvenor, Andrew, 101 Gudat, Wolfgang, 26, 28 Gupta, Raj, 29

Hackett, Peter, 80 Hallin, Emil, 39, 40f, 50, 83–4, 96–7, 102–3, 150–1 Halliwell, Janet, 46 Hansen, Brian, 48–9, 51, 63–9, 71–2, 74–80, 82–3, 93 Hansson, Carolynn, 56 Harding, Paul, 53 Harrington, Ertle, 4, 9–10 Harris, Mike (prem. Ontario), 53 Hartman, Francis, 84 Harvey, Tracene, 139 Haslam, R.N., 10 Hawkins, Barry, 59, 84, 96, 101, 150 Haworth, Richard, 124 Hawthorn, David, 129 Hayden, Don, 53, 57 Hayden, Michael, 11–12, 19, 66, 119 He, Feizhou, 129 health research, 6, 101, 119, 126, 130–2; Alzheimer’s disease, 131–2; cystic fibrosis, 126, 132; protein structures, 130–1 Heikoop, Martin, 96, 101 Henderson, Grant, 40, 102 Hermes, Joseph, 88 Herzberg, Gerhard, 4, 8–9, 22, 24; Atomic Spectra and Atomic Structure, 9; Luise (spouse), 8 Hitchcock, Adam, 6n, 31, 35, 38–40, 39f, 43, 53–5, 57, 60, 72, 76, 82, 89–90, 95, 97, 131, 150–1 Hochst, Hartmut, 122 Horachek, Jane, 64, 77 Hormes, Josef, 94, 137, 139 Hosek, Chaviva, 50 Hu, Yongfeng, 38, 40f, 102–3, 127, 129 Hyshka, John, 49, 59 Igloo, 16 Industrial Strategy Committee, 53

Index 163 Innovation Place, 48, 58–9, 63–4, 66, 68–9 Interface Science Western (ISW), 25, 57, 61 International Vaccine Centre (INTERVAC), 105, 114t Irwin, Marty, 51 Ivany, George, 19, 46–50, 55, 58, 63–6, 68, 70–1, 74–5, 77, 80, 82–3, 85, 120, 150 Jacobs, Pat, 25 Jiang, De-Tong, 38–9, 97, 102–3 Johns, Harold, 4, 9–10 Johnson, Dennis, 2, 19–20, 20f, 47–51, 54, 58–60, 63–5, 67–73, 75–7, 81–2, 84, 90, 92, 97, 101–2, 150, 162–3 Jong, Mark de, 88, 93–9, 101, 137–8 Kang, Yong, 53, 56 Karunakaran, Chitra, 133, 135 Kasrai, Masoud, 29, 35–6, 38–40, 39f, 52, 87, 122–5; Guiti (spouse), 124–5; Leila (daughter), 124–5; Reza (son), 124–5 Katz, Leon, 4, 9–11, 14–15, 14f Kaznatcheev, Konstantine, 103 Kendrew, John, 130 Kennedy, Laura, 80 Kitchen Cabinet committee, 48–9, 54–5, 59, 63, 66, 68, 79 Kline, Gerry, 3, 11n Kobilka, Brian, 130 Kozachuk, Madalena, 139 Kratchovil, Ron, 90 Kremers, Onno, 75–6, 150–1 Laclare, Jean-Louis, 84 Lamb, Rob, 88, 104, 125, 138, 141 Lapeyre, Gerry, 26–7; G.J. Lapeyre award, 27, 34 Lapointe, Paul-Henri, 71 Lau, Leo, 29

Lazier, Bob, 24 Lennard, Willy, 29 Lepage, Beryl, 88–92, 95–6, 102, 114–15, 138 Lepage, Marc, 51 Lien, Neil, 32 Lindau, Ingolf, 55–6 Lindley, Peter, 84–5 linear accelerators (LINACs), 11–14, 14f, 16, 24, 59, 137–8, 140–1, 143 Lingenfelter, Dwain (MPP), 64, 74–5 Longstaffe, Fred, 87 Lorimer, Jack, 31 Lynch, Kevin, 50 Macdonald, Donald (MP), 50, 70 MacKenzie, C.J., 10 MacKinnon, Janice (MPP), 50, 80, 150 MacKinnon, Peter, 5, 13, 47, 52, 88, 94–5, 98–103, 103f, 115, 117–20; University Leadership and Public Policy in the Twenty-First Century, 52, 66, 98–9, 118, 120 Maddock, A.G., 22, 124 Madey, John, 56 Maley, Doug, 75, 78, 80, 84, 91, 95, 99, 150–1 Manley, John (MP), 67, 70–4, 80 Mansbridge, Peter, 103 Marchi, Sergio (MP), 67 Martin, Paul (MP, PM), 5, 44, 47, 50, 67, 70–4, 82, 106, 121, 151 Martin, Ron, 139 Masse, Marcel, 67, 70 MAX IV (Sweden), 140–1; BioMAX protein crystallography beamline, 141; FemtoMAX beamline, 141 May, Tim, 39, 40f, 102–3 Mays, Martin, 23 McAlpine, Bob, 45–7, 56, 76, 151 McAuley, Alex, 55, 58, 66, 76, 77, 96, 151; McAuley report, 60, 65, 69–70

164 Index McCann, Peter, 51, 59–60 McCrindle, Bob, 56 McDonald, Art, 115 McFarland, Jim, 75 McGowan, Bill, 4, 16, 23–5, 28–30, 32, 33f McIntyre, Stewart, 24, 29, 53, 57 McKonkey, Bill, 31 McLaughlin, Murray, 48, 68 McLellan, Anne (MP), 67 McNabb, Gordon, 30 McPherson Engineering, 31–2, 37–8 McPherson ESCA 36 spectrometer, 23–4 Medical Research Council (MRC), 51, 71, 90, 100–1, 107–8, 151, 162. See also Canadian Institutes of Health Research (CIHR) Mercer, Peter, 55 Messer, Jack, 59 Michel, Bernard, 70, 75, 80, 82, 150 Middleton, Fred, 35 Miller, David, 50 Mills, Al, 56 Mills, Dennis (MP), 2, 50–1 Milton, Stephen, 56, 85 Mitchell, Ian, 29 Moewes, Alex, 100 Morand, Peter, 44–5, 151 Mössbauer spectroscopy, 22–3, 161; Rudolph Mössbauer, 22 Muller, Alex, 128 multi-bend achromat (MBA) storage ring, 140–1 Murray, Walter, 118 Najman, Morey, 36 National Blueprint Document, 66–7 National Research Council (NRC), 9–11, 24–5, 28–34, 41, 76–8, 80–6, 95, 99, 107, 115, 149–51, 162; funding

support, 4, 10–11, 15, 21–2, 24–5, 29–30, 33–4, 36–8, 83, 84t, 109; Plant Biotechnology Institute, 58 National Research Universal (NRU) reactor, 45 National Sciences and Engineering Research Council (NSERC), 5–6, 15–18, 41–50, 52, 55–6, 66–7, 69–71, 81–5, 139, 149, 151; Collaborative Research Initiatives Program, 43; Committee on Materials Research Facilities (CMRF), 44, 46; funding support, 15, 18–19, 21–2, 29–31, 33– 4, 36–8, 40, 43, 45–8, 61, 63, 101–2, 108–9; Site Review Committee, 56–61, 85, 149 National Synchrotron Light Source (NSLS), 56, 133, 135 Natural Resources Canada (NRCan), 83, 84t, 92, 133 Nelson, Andrew, 139 Nesbitt, Wayne, 24 Neville, John, 38 Nicholls, Mark, 36 Nicholson, Fraser, 78, 80, 82, 95, 150 Norton, Peter, 29 Norum, B.E., 16, 60 Nystuen, Gordon, 71 Olsson, Bengt, 35–6 Ontario Centre for Materials Research, 35 Ontario Innovation Trust (OIT), 87, 89–91, 108; Agreement, 91 Ontario Synchrotron Consortium (OSC), 90–1; Agreement, 91 Pai, Emil, 57 Park, Anne, 71 Peak, Derek, 101 Pedersen, George, 51

Index 165 Pensa, Claude, 51 People’s Free University, 119 Perutz, Max, 130 Petersen, Nils, 37f, 95 Physical Sciences Laboratory (PSL), 25, 32, 35 Pickering, Ingrid, 100 Pink, David, 55, 66 Pope, Tim, 53 Project Soleil (France), 84, 149 PromoScience, 139 Protein Data Bank, 126, 130 Pruett, Charlie, 26–7, 29, 32, 122 Pywell, Rob, 102 Quail, Wilson, 44, 98 Quantum Materials Program, 129 Quantum Matter Institute, 129 Redhead, Paul, 28 Regier, Tom, 129, 138 Retzlaff, Greg, 151 Ribeiro, Sandra, 138 Richardson, Doug, 19, 20f, 44, 47–50, 59–61, 63–6, 68, 70–1, 74, 77, 82–3, 92, 101–2, 150 Robayo, Louis, 78 Robinson, Paul, 56 Romanow, Roy (prem. Sask.), 47–8, 50, 59–60, 69, 150 Rowe, Ed, 16, 24–7, 29, 33f, 34, 36, 41, 122 Runte, Roseann O’Reilly, 109 Sandorfy, Camille, 28 Saskatchewan Accelerator Laboratory (SAL), 11–21, 13f, 14f, 43–4, 48, 52, 56–61, 66–9, 73–4, 78, 80, 83–4, 86, 92, 102, 137, 139, 143, 146, 150; funding, 11n, 12, 15, 19, 24, 29, 43–4, 46, 63

Saskatchewan Centre for Synchrotron Research, 99 Saskatchewan Opportunities Corporation, 68, 163 Saskatchewan Synchrotron Institute (SSI), 92, 99–101, 163 Saunderson, William (MPP), 58 Sawatzky, George, 129 scanning transmission x-ray microscopy (STXM), 131, 133–4, 141 Second Century Fund, 30–1 SED Systems, 58 Servranckx, Roger, 15–17, 25, 60; pulse stretcher ring (PSR), 15 Sham, T.K., 25–6, 29, 31, 35–6, 37f, 38, 39f, 40, 43, 52–3, 57, 87, 90, 102, 122, 127, 139, 150; Soochow University-Western University Centre for Synchrotron Radiation Research, 127 Shapiro, Bernard, 70 Shenher, Angela, 102 Sheppard, Malcolm, 48–9 Sheridan, Georgette (MP), 64 Sherman, Norman, 31, 33–4, 33f, 151 Shinozaki, Doug, 32 Shoesmith, Dave, 29 Siegbahn, Kai, 24, 28 Siegbahn group (Upsala University), 24 Silzer, Mark, 17, 17f, 60 Skopik, Dennis, 2, 14, 16, 18–20, 18f, 31, 43, 46–51, 59–60, 63–5, 67–72, 75–8, 80–4, 86, 95, 102, 146, 149–50 Slinger, Rob, 92, 99 Smith, Bill, 78, 99 Smith, Ian, 22 Smith, Lorne, 48–9 Social Science and Humanities Research Council (SSHRC), 100, 108 Spannier, Larry, 75, 78, 80, 82–4, 91, 93, 95, 99, 150

166 Index spherical grating monochromator (SGM), 37–8, 37f, 40, 129, 135 Spinks, John, 8–9, 11–12, 13f SPring-8 Angstrom Compact freeelectron laser (Japan), 140 Stakiw, Chris, 66 Stanford synchrotron, 16n, 25, 42, 100, 140, 148 Stanford Synchrotron Institute, 100 Stenhouse, Ian, 124 Stiller, Cal, 90 Stillman, Martin, 53 Stoicheff, Boris, 30, 46 Strangway, David, 82 Strom-Olsen, John, 151 Sudbury Neutrino Observatory, 115 Sulzenko, Andrei, 71 Sun, Andy, 127–8; Advanced Materials for Clean Energy, 127 Surface Science Western (SSW), 25, 29–31, 45, 53, 55, 57, 61, 89, 161; funding, 29–31 Sutarto, Ronny, 129 Svensson, Svente, 28 Swiss Light Source, 131 Synchrotron Radiation Center (SRC), 21–2, 25, 27, 34, 36, 38, 41, 122 Synchrotron Steering Committee. See Kitchen Cabinet committee Tan, Kim, 27, 31–4, 33f, 37f, 38, 40f, 102, 122, 139, 151 Tantalus synchrotron, 16, 21, 24–7, 29, 32–4, 33f, 146 Tastad, Doug, 48, 59 Taylor, Allan, 68–70 Taylor, Jim, 37f Tchorzewski, Ed, 50 Thomas, John, 23–4 Thomlinson, Bill, 88, 94, 97–8, 103 Thompson, James S., 10

Thomson, J.J., 147 Tiedje, Tom, 60 Traum, Mort, 27 Tri-University Meson Facility (TRIUMF), 12, 43, 47, 59, 77, 115 Trudeau, Pierre, 70 Trump, Donald (pres. US), 35, 123 Tse, John, 39, 96–7, 100, 102 Turner group (Oxford University), 24 Tylisczczak, Tolek, 35 U15, 116–17 Underwood, McLellan and Associates (UMA), 14, 44, 59–60, 78, 80, 92, 96, 101, 150 Urquhart, Stephen, 39 USask Steering Committee, 77–8, 80, 83 USask synchrotron proposal, 19, 47–9, 52, 54–5, 63, 66, 68, 79; cost estimates, 59, 61; joint Western proposal, 63–4, 69; NSERC site visit, 58–60 USask-CLS agreement, 93–4 Vaccine and Infectious Disease Organization (VIDO), 6, 58, 105–6, 106t, 114t, 117, 119 Walker, Tracy, 138–9 Walzak, Mary-Jane, 53 Wang, Jian, 129, 131, 134 Ware, Bill, 29–30 Webster, Matt, 84, 90–2 Western Economic Diversification (WED), 44, 46, 58, 60, 61, 63, 75, 78, 83–4, 84t, 91–2, 95, 99, 109, 119 Western University synchrotron proposals, 19, 28, 37, 43, 47–9, 52–5, 63, 66, 68, 79; cost estimates, 57, 59, 61; Development/Management Committee, 53; joint USask proposal, 63–4, 69; NSERC site visit, 55–60

Index 167 Whitworth, Tony, 49, 63, 68, 71–2, 75, 77, 80, 87, 93–6, 150 Wiens, Ed, 84, 99 Williams, Gwyn, 56 Willis, Clive, 76–8, 82, 84, 150 Winston, John, 53–4, 56, 58 Woodhouse, Harold, 136 Woodward, Ron, 48 Wright, John, 76, 78 Wyant, Peter, 76–8, 84, 150 Wyatt, Hal, 48, 75–7, 80, 150

x-ray free electron laser (XFEL), 140–1 Yang, Bing Xing, 35–6, 122, 151 Yates, Brian, 32–4, 33f, 37–8, 37f, 102–3, 151 Yin, Zhangfeng, 36 Yip, Chris, 53, 56 Youzwa, Pat, 50, 59 Yuel, Jim, 63–4, 66, 68–9, 77 Zhou, Jigang, 40f, 102, 129

About the Authors

G. Michael Bancroft

G. Michael Bancroft’s paternal grandparents homesteaded in Arcola, Saskatchewan, in 1903. His father grew up in Regina and eventually worked in Winnipeg and Saskatoon, where Bancroft was born in 1942. He received his BSc (Honours) and MSc from the University of Manitoba and his PhD from the ­University of Cambridge in 1967. That year he became a research fellow at Christ’s College, Cambridge, and a demonstrator at the University of ­Cambridge, a position he held until 1970, when he joined Western’s Department of Chemistry as an a­ ssistant professor in 1970. He quickly rose through the ranks, becoming a full professor in 1974. He served as department chair from 1986 to 1991 and again from 1992 to 1995. He was director of the Centre for Chemical Physics from 1978 to 1981, and in those four years he established two internationally recognized facilities: ­Surface Science Western and the Canadian Synchrotron Radiation Facility at the University of Wisconsin-Madison. Bancroft is best known as a pioneer in the use of Mössbauer spectroscopy, photoelectron spectroscopy, and synchrotron radiation to record high-resolution spectra of minerals, glasses, inorganic molecules, coals, and surfaces. He has published over 430 papers and supervised 39 graduate ­students and 36 post-doctoral fellows. Known as a leading force in Western’s transformation into a research-intensive university, Bancroft was also a Canadian leader in the establishment of the ­Canadian Light Source project that went to Saskatoon. He has been recognized for his scholarly work with numerous prizes, ­including Western’s Florence Bucke and Hellmuth Research Awards, and many others from the Royal Society of Canada (of which he is a fellow), the Canadian Society of Chemistry (the CIC Medal and the Montreal Medal), the ­American Chemical Society, and the Chemical Society of Great Britain. He has been

170  About the Authors

awarded three honorary degrees: from Western, the University of Manitoba, and St. John’s College at the University of Manitoba. His multi-decade-long dream of establishing a national synchrotron facility became a reality when the Canadian Light Source was built at USask. He served as the facility’s first director from 1999 to 2001, and was its acting research director from 2001 to 2005. For this achievement, he became an officer of the Order of Canada in 2003. But more importantly, he has helped provide the infrastructure and inspiration for future generations of Canadian scientists working in this important field. He has been married to Joan for over fifty years, has two children, David (Michelle) and Catherine, and four grandchildren, Ethan, Avery, Kenneth, and William. Dennis D. Johnson

Dennis Johnson was born in Saskatoon and raised in LeRoy, a farming community in rural Saskatchewan. He graduated from the University of Saskatchewan with a BSc in pharmacy (1960) and a MSc in pharmacology (1962). He received his PhD in pharmacology from the University of Washington (Seattle) before returning to Canada to take up an appointment in the Department of Physiology and Pharmacology at the University of Saskatchewan in 1965. Dr. Johnson has over forty years of experience in research and development. He has held research grants throughout his career and for fifteen years served at an administrative level. From 1965 to 1992 he taught pharmacology in the ­College of Medicine and conducted research in neuropharmacology. He served as head of the Department of Pharmacology and as assistant dean (research) in the ­College of Medicine. From 1992 to 1997 he served as the associate vice-president (research) for the University of Saskatchewan. He retired from the university in 1997 but continued to work as a consultant in science and technology. Dr. Johnson served for two terms as a council member and on the executive of the Medical Research Council of Canada. He was on the executive of the Canadian Federation of Biological Societies, president of the ­Pharmacological Society of Canada, chair of the Saskatchewan Health Research Board, and on the board of Epilepsy Canada. He served on the advisory boards of the Pharmaceutical Manufacturers Association of Canada Health Research ­ ­Foundation, the National Cancer Institute of Canada, the Plant Biotechnology Institute of the National Research Council, and as a member of the National Biotechnology Advisory Committee of the Government of Canada. While serving as associate vice-president (research) at USask he was on the Board of Directors of the ­Saskatchewan Research Council and the Vaccine and ­Infectious Diseases Organization, and chair of the Saskatchewan Drug Research Institute and the University of Saskatchewan Technologies Inc.

About the Authors  171

He is a past president of the Saskatoon Chamber and was on the Board of Directors of the Saskatchewan Opportunities Corporation, the Saskatchewan Agrivision Corporation, and the Canadian Ukraine Centre. He was chair the Meewasin Valley Authority Board of Directors from 2003 to 2005. Dr. Johnson was involved in the creation of the Saskatchewan Drug ­Research Institute and the Cameco MS Neuroscience Research Centre, and he was ­instrumental in recruiting the Canadian Light Source to the University of ­Saskatchewan. He was director of the Saskatchewan Synchrotron Institute from 2002–4. He was awarded a Rotary Golden Wheel Award for contributions to the city of Saskatoon and on two occasions was recognized as a Saskatonian who “made a difference.” In 2014, he received a Centennial Alumni of Influence Award from the College of Pharmacy and Nutrition. Dennis is married to Sharon. They have four children and nine grandchildren.

Jacket Illustrations

Front top left: A close-up of one of the CLS experimental chambers for detailed analysis of materials (such as advanced solar cells using x-ray scattering) using the CLS high-energy x-ray beamline. The sample for analysis is mounted in the middle of the chamber, with the laser-like x-ray beam from the synchrotron beamline impinging on the sample from the right. Canadian Light Source Inc. Front bottom: A view of part of the CLS facility taken from the mezzanine. The top right half of the image shows the circular booster electron injector shielded by thick white concrete. Beige power supplies are seen inside the booster ring. The lower right of the image shows a small arc of the main synchrotron, which emits the usable radiation, with racks of shielded wires on top. Canadian Light Source Inc. Back: An arc of the synchrotron with conduit trays filled with power and signal wiring, cooling water pipes, and the large air-handling pipe on top of the synchrotron. In the lower left corner, the roof of the mid-infrared end-station lab is visible. Canadian Light Source Inc.