Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Critical Materials
Critical Materials
Critical Materials
Ebook559 pages4 hours

Critical Materials

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Critical Materials takes a case-study approach, describing materials supply-chain failures from the bronze age to present day. It looks at why these failures occurred, what the consequences were, and how they were resolved. It identifies key lessons to guide responses to current and anticipated materials shortages at a time when the world’s growing middle class is creating unprecedented demand for manufactured products and the increasingly exotic materials that go into them. This book serves as a guide to materials researchers and industrial end-users for finding effective approaches to shortages of specialty materials.

The lessons in the book are also appropriate to those who use materials and for those involved in manufacturing supply-chain management and industrial design.

  • Instructs the reader on how to select the most effective strategies to deal with materials supply-chain failures
  • Discusses technical feasibility, economic viability and the political context
  • Includes an extensive use of case studies to illustrate key concepts of criticality
LanguageEnglish
Release dateOct 22, 2020
ISBN9780128187906
Critical Materials
Author

Alexander King

Alex King is a professor of Materials Science and Engineering at Iowa State University. He earned his doctorate from Oxford University and did his post-doc work at both Oxford University and MIT. He went on to join the faculty at the State University of New York at Stony Brook, where he also served as the Vice Provost for Graduate Studies (Dean of the Graduate School). He was the Head of the School of Materials Engineering at Purdue from 1999 to 2007. From 2008 until 2013 he was the Director of DOE’s Ames Laboratory and became the Founding Director of the Critical Materials Institute from 2013 through 2018. Dr. King is a Fellow of the Institute of Mining Minerals and Materials; ASM International; and the Materials Research Society. He was also a Visiting Fellow of the Japan Society for the Promotion of Science in 1996 and a US Department of State Jefferson Science Fellow for 2005-06. He served as the President of MRS in 2002, Chair of the University Materials Council of North America from 2006-07, Co-chair of the Gordon Conference on Physical Metallurgy in 2006, and Chair of the APS Interest Group on Energy Research and Applications for 2010. Dr. King was named the recipient of the 2019 Acta Materialia Hollomon Award for Materials and Society. Alex King delivered a TEDx talk on critical materials in 2013 and was the TMS & ASM Distinguished Lecturer on Materials and Society in 2017. He is currently a scientific adviser for Harvard’s Material Alchemy (described as “translating science into commercial products that use sustainable materials”) and a member of the Advisory Board of CHiMaD (the Center for Hierarchical Materials Design, funded by the Department of Commerce, and led by Northwestern University).

Read more from Alexander King

Related to Critical Materials

Related ebooks

Materials Science For You

View More

Related articles

Reviews for Critical Materials

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Critical Materials - Alexander King

    Critical Materials

    First Edition

    Alexander King

    Table of Contents

    Cover image

    Title page

    Copyright

    Acknowledgments

    Abbreviations and acronyms

    Preface

    1: What happened to the rare earths? Monopoly, price shock, and the idea of a critical material

    Abstract

    The essentiality of the rare earths

    Rare earth sources

    Rare earth supply challenges—2005–15

    After the price spike

    Lessons learned: The price spike in hindsight

    Coda: Every challenge is also an opportunity

    2: This is not new. A short history of materials criticality and supply-chain challenges

    Abstract

    Copper and the end of the Bronze Age (~ 1200 BCE)

    The Venetian monopoly on glass

    Cordite in World War I (1914–18)

    Silk and nylon (the 1930s and 1940s)

    World War II (1939–45)

    Old lead (1978)

    Cobalt (1978)

    Niobium (1979)

    Molybdenum (1980 and 2004)

    Tantalum (1997, 2000, and 2008)

    Photovoltaic silicon (mid-2000s)

    Rhenium (2006–08)

    Lessons learned

    3: Assessing the risks

    Abstract

    Defining critical materials

    Assessments of materials criticality

    Consistency and contrast

    Variation of criticality over time

    Regional perspectives on criticality

    Indicators of criticality

    What does criticality mean?

    Consequences of criticality

    Tipping points. What takes us from criticality to crisis?

    Lessons learned

    What will we need?

    4: What changed after the rare earth crisis?

    Abstract

    Impacts of the rare earth crisis

    Conflicts and conflict resolution

    The supply side

    The demand side

    Postcrisis rare earth prices and utilization

    Lessons learned

    5: Mitigating criticality, part I: Material substitution

    Abstract

    The challenge of inventing materials on demand

    Improving the forecast

    Using existing materials

    Increasing the speed of new material discovery and deployment

    Lessons learned

    6: Mitigating criticality, part II: Source diversification

    Abstract

    How are mines developed?

    Conventional mines

    Unconventional sources

    Coproduction

    Progress since the rare earth crisis

    Lessons learned

    7: Mitigating criticality, part III: Improving the stewardship of existing supplies

    Abstract

    Urban mines versus conventional mines

    Regulatory versus economic drivers

    Reducing manufacturing waste

    In-process recycling

    End-of-life recycling

    Recycling as a response to criticality: Successes and failures

    What fraction of current need can be met by recycling?

    Emerging targets for recycling

    Potentially viable recycling technologies

    Lessons learned

    8: Tactics and strategies for the future

    Abstract

    What have we learned?

    Time is the biggest challenge

    Summary

    Epilog: Criticality in the time of coronavirus

    Index

    Copyright

    Elsevier

    Radarweg 29, PO Box 211, 1000 AE Amsterdam, Netherlands

    The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, United Kingdom

    50 Hampshire Street, 5th Floor, Cambridge, MA 02139, United States

    © 2021 Elsevier Inc. All rights reserved.

    No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopying, recording, or any information storage and retrieval system, without permission in writing from the publisher. Details on how to seek permission, further information about the Publisher’s permissions policies and our arrangements with organizations such as the Copyright Clearance Center and the Copyright Licensing Agency, can be found at our website: www.elsevier.com/permissions.

    This book and the individual contributions contained in it are protected under copyright by the Publisher (other than as may be noted herein).

    Notices

    Knowledge and best practice in this field are constantly changing. As new research and experience broaden our understanding, changes in research methods, professional practices, or medical treatment may become necessary.

    Practitioners and researchers must always rely on their own experience and knowledge in evaluating and using any information, methods, compounds, or experiments described herein. In using such information or methods they should be mindful of their own safety and the safety of others, including parties for whom they have a professional responsibility.

    To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors, assume any liability for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions, or ideas contained in the material herein.

    Library of Congress Cataloging-in-Publication Data

    A catalog record for this book is available from the Library of Congress

    British Library Cataloguing-in-Publication Data

    A catalogue record for this book is available from the British Library

    ISBN: 978-0-12-818789-0

    For information on all Elsevier publications visit our website at https://www.elsevier.com/books-and-journals

    Image Credit: Trevor M. Riedemann. This photograph has been authored, in whole or in part, under Contract No. DE-AC02-07CH11358 with the U.S. Department of Energy.

    Publisher: Matthew Deans

    Acquisitions Editor: Christina Gifford

    Editorial Project Manager: Isabella C. Silva

    Production Project Manager: Vijayaraj Purushothaman

    Cover Designer: Miles Hitchen

    Typeset by SPi Global, India

    Acknowledgments

    This book is an attempt to summarize what I have learned in the process of setting up and running the Critical Materials Institute for its first 5 years, and I have had many teachers. The late Karl Gschneidner, Jr. was my mentor on all things related to the rare-earth elements, and he is greatly missed. All of the people of CMI have contributed to my education but especially its leadership team, which has included Iver Anderson, Gretchen Baier, Joni Barnes, Deb Covey, Rod Eggert, Cynthia Feller, Yoshiko Fujita, Dan Ginosar, Chris Haase, Carol Handwerker, the late Scott Herbst, Ed Jones, Tom Lograsso, Scott McCall, Bruce Moyer, Eric Peterson, Brian Sales, Adam Schwartz and Eric Schwegler. Others have provided notable input, specific contributions, and moments of enlightenment or levity, including Jenni Brockpahler, Steve Constantinides, Bill McCallum, Ikenna Nlebedim, Duane Johnson, Richard LeSar, John Ormerod, Ryan Ott, Orlando Rios, Sadas Sadasivan, Alok Srivastava, Stan Trout, Patrice Turchi, Jeff Wang, David Weiss, and probably a few others who have slipped my memory.

    The US Department of Energy supported the learning phase of this work by funding the Critical Materials Institute through the Division of Energy Efficiency and Renewable Energy’s Advanced Manufacturing Office. This book was written with the support of the Iowa State University of Science and Technology.

    Finally, my wife, Christine, has provided steadfast forbearance and stability amid the chaos, along with some timely professional consultation on information resources.

    Abbreviations and acronyms

    ACREI Association of Chinese Rare Earth Industries

    AI artificial intelligence

    AIM accelerated insertion of materials

    AIST The National Institute of Advanced Industrial Science and Technology (Japan)

    ALMR Association of Lamp and Mercury Recyclers

    AMD acid mine drainage

    AMR anisotropic magnetoresistance

    ASTM American Society for Testing and Materials

    ASX Australian Stock Exchange

    At. No. atomic number

    BCE before the Christian (or common) era

    BFR brominated flame retardant

    BGS British Geological Survey

    CAGR compound annual growth rate

    CALPHAD calculation of phase diagrams

    CAPEX capital expenditures

    CENRS Committee on Environment, Natural Resources, and Sustainability (of NSTC)

    CES Cambridge Engineering Selector (from Granta Design)

    CFL compact fluorescent lamp

    CMI Critical Materials Institute

    COVID-19 coronavirus disease 2019

    CPU central processor unit

    CRIRSCO Committee for Mineral Reserves International Reporting Standards

    CRM critical raw material

    CRT cathode ray tube

    DARPA Defense Advanced Research Projects Agency (the United States)

    DfD design for disassembly

    DFS definitive feasibility study

    DFT density functional theory

    DHS Department of Homeland Security of the United States

    DKB designer knowledge base

    DOE Department of Energy of the United States

    DRAM dynamic random-access memory

    DRC Democratic Republic of the Congo (since 1997), formerly Zaire

    EC European Commission

    EEZ exclusive economic zone

    EU European Union

    EV electric vehicle

    FOB free on board at shipping point (means that ownership is transferred to the purchaser when goods leave the named shipping point)

    GM General Motors

    GMR giant magnetoresistance

    GPS global positioning system

    GRS government rubber styrene

    GTP Global Tungsten and Powders Corporation

    HD hydrogen decrepitation

    HDD hard disk drive

    HDDR hydrogenation, disproportionation, desorption, and recombination

    HEU high-enriched uranium

    HEV hybrid electric vehicle

    HREE heavy rare-earth element

    HSLA high strength, low alloy (of steels)

    IC integrated circuit

    ICE internal combustion engine

    ICME integrated computational materials engineering

    iNEMI International Electronics Manufacturing Initiative

    IR infrared

    ISO International Standards Organization

    JCG Japan Coast Guard

    JORC Joint Ore Reserves Committee (produces the Australasian code for reporting of exploration results, mineral resources, and ore reserves)

    LCO lithium cobalt oxide

    LENS laser engineered net shaping

    LEU low-enriched uranium

    LIBS laser-induced breakdown spectroscopy

    LME London Metal Exchange

    LREE light rare-earth element

    MD molecular dynamics

    MESS multiple elements from a single source

    MGI materials genome initiative

    MRI magnetic resonance imaging

    MRT molecular recognition technology

    MSX membrane solvent extraction

    NASA National Aeronautics and Space Administration

    NEDO New Energy and Industrial Technology Development Organization (Japan)

    NI-43101 National Instrument 43-101 (Canadian resource assessment and reporting standard)

    NIST National Institute of Standards and Technology (the United States)

    NMC lithium nickel manganese cobalt oxide

    NORM naturally occurring radioactive material

    NRC National Research Council (of the US National Academies of Sciences, Engineering, and Medicine)

    NSTC National Science and Technology Council (of the United States)

    NYMEX New York Mercantile Exchange

    OPEX operating expenditures

    PEA preliminary economic assessment

    PFS preliminary feasibility study

    PGM platinum group metal

    ppm parts per million

    PRC Peoples’ Republic of China

    PV photovoltaic

    PVC polyvinyl chloride

    QP quench and partition

    R&D research and development

    RAM random-access memory

    REE rare-earth element

    REO rare-earth oxide

    REPM rare-earth permanent magnet

    REY rare earths plus yttrium, i.e., the lanthanides plus yttrium

    ROI return on investment

    ROW Rest of World, i.e., all countries except China

    SAMREC South African Mineral Reporting Codes

    SCSMSC Subcommittee on Critical and Strategic Mineral Supply Chains (of NSTC)

    SOFC solid oxide fuel cell

    SX solvent extraction

    TBC thermal barrier coating

    TREO total rare-earth oxide, i.e., the sum over all of the rare-earth elements

    TRL technology readiness level

    TSX Toronto Stock Exchange

    US, USA The United States (of America)

    UV ultraviolet

    WTO World Trade Organization

    WWI World War I

    WWII World War II

    XRF X-ray fluorescence

    YSZ yttria-stabilized zirconia

    Preface

    Alexander King, Ames, IA, United States

    I did not set out to write this book.

    In 2010 a group of researchers gathered to design an agenda for research and development that aimed at alleviating the challenges to rare-earth supplies that were then emerging. In 2013 the Critical Materials Institute was stood up, and I served as the director for its first 5 years. As we and others around the world pursued various lines of research, we learned a lot of things that did not work and also found a few that did. This volume is an attempt to summarize that knowledge, as logically and methodically as such recent hindsight allows, to create a guide for others working threats to materials supply chains.

    This is not a book about materials science.

    Although a certain amount of materials science is included, it is usually only described at the level of an introductory college class, and much more detailed texts are available if you want to read deeply into that subject. This book is about the intersections of materials science, manufacturing and public policy. My goal has been to include just enough about each of these topics to enable the practitioners of the others to understand each other and have enough vision into their different worlds to recognize where their efforts may or may not align. When they align, impact is possible; when they do not, all effort is for naught.

    I draw on historical and recent examples of R&D efforts that have helped to alleviate material shortages, along with cases where they have failed to have any impact despite the excellence of the research. Lessons learned from these case studies are used to identify the traits of successful programs and the pitfalls that should be avoided in creating research-based solutions to material supply challenges.

    1: What happened to the rare earths? Monopoly, price shock, and the idea of a critical material

    Abstract

    Rare earth prices surged to unprecedented levels in 2011 creating concerns about the security of their supplies and triggering a broad range of responses. Some of the conditions that led to the rare earth crisis also apply to other industrial feedstocks and the interrelated economics and technologies of critical materials rapidly emerged as an urgent area of research and development activity.

    Keywords

    Rare earths; Price shock; Monopoly; China; Critical material

    Every manufactured thing is made of some kind of material.

    In some cases manufacturers are free to choose among different materials that can meet their needs, and they make selections according to the cost or properties of the available materials, to optimize profit or performance.

    In other cases there is only one material that meets the needs of the manufactured object. Thomas Edison’s incandescent electric lightbulb famously depended on the identification of a material to use for the hot, glowing filament, and after testing hundreds or maybe thousands of possible materials, tungsten was identified as the only one that met the needs: It could be heated by an electric current, and it did not melt at the elevated temperatures where it would glow and provide light.

    Materials that are necessary for a particular product, like tungsten in lightbulbs, are regarded as essential. You can’t make the product without them, and there are no ready substitutes to use if you can’t get them. Today’s technologies are full of essential materials: All of our smart devices rely on data processing enabled by silicon and several ancillary materials; nuclear reactors need uranium to produce heat that is eventually converted to electricity; early photocopiers and laser printers were enabled by the particular properties of selenium sulfide print drums; electrical energy storage is dominated by rechargeable batteries made with lithium; and despite increasing use of carbon fibers, the aviation industry still depends very heavily on aluminum-based structures.

    Essential materials have particular properties that make them work: Tungsten has a very high melting point, silicon is a semiconductor, uranium’s atomic nucleus can be fissile, selenium sulfide is a photoconductor that only transmits electricity when it is exposed to light, lithium is easily ionized and transported between the electrodes of batteries, and aluminum has high strength relative to its weight and cost. Because of their properties, these (and many other) materials are, or have been, essential. As will be clear from some of these examples, essentiality is not a permanent property of a material: Incandescent lighting has been replaced by other more efficient technologies, nuclear power is increasingly unpopular in many parts of the world, selenium sulfide has been replaced by amorphous silicon in laser printers, and aluminum is being replaced by carbon fiber composites in some airplanes.

    The rare earths are a particular set of chemical elements whose properties make them essential to a wide range of products. These once-obscure substances started to gain attention in the late 2000s because they are essential to many technologies, and their supplies were perceived to be threatened.

    When an essential material has a risk of supply disruptions, it is considered a critical material, and when supply concerns erupted about the rare earths, they sparked broader interest in the general issue of critical materials. A decade after the rare earth price shock, we have learned a great deal about criticality, materials supply crises, and how best to manage them.

    The essentiality of the rare earths

    The rare earths comprise 17 chemical elements including all of the lanthanides, the 15 elements from lanthanum (atomic number 57) to lutetium (At. No. 71). Scandium (At. No. 21) and yttrium (At. No. 39) are usually included in the set of rare earth elements (REEs), because they also belong to Group 3 of the periodic table and are closely related to the rare earths in terms of their chemical behavior. All of the rare earth elements are shown in Fig. 1.1.

    Fig. 1.1 The rare earth elements include all of the lanthanide series along with scandium and yttrium.

    The definition of the set of elements that belong to the rare earth category is not always consistent. Some authors exclude or ignore scandium, and some authors refer to the rare earths plus yttrium (REY) implying that they only consider the lanthanides to be true rare earths. Distinctions are frequently drawn between light and heavy rare earths, based upon their atomic weights, but the borderline between light and heavy is variable, and some people apply the label medium to an intermediate group, also of variable membership. Despite this variability, distinctions between light and heavy rare earths are useful. Rare earth mines are typically described as being richer or poorer in one or other category—and yttrium is more often associated with heavy rare earths in this context, so it is considered a heavy rare earth, despite its relatively low atomic weight.

    The first rare earths were discovered in 1787, in the form of a black mineral from a quarry in Ytterby, Sweden, near Gothenburg. From this rock the Finnish chemist Johan Gadolin extracted an oxide (or earth), which he named ceria, after Ceres, the Roman goddess of agriculture, grain crops, fertility, and motherly relationships. He published the discovery in 1794 [1]. Only later were the various rare earths separated from each other and reduced into elemental metals, but as they were successively isolated, uses for them also developed. A newly discovered rare earth bearing mineral was named gadolinite in 1800, in Gadolin’s honor. When element number 64 was discovered in 1880, it was named after him, too—gadolinium—even though it is not extracted from gadolinite.

    Like gadolinium, most of the rare earths were discovered in the late 19th and early 20th centuries, following the development of the periodic table by Mendeleev, in 1867 [2]. The systematic organization of the chemical elements predicted the existence of elements that were not yet known, and a large part of chemical science of that time was devoted to finding them. It was assumed that the lanthanides must be rare because they were not found very easily, and the term rare earth stuck. In fact the rare earth elements are not so much rare as they are elusive. Except for promethium (At. No. 61), which decays radioactively, has no stable isotopes, and is not found in nature, the rare earths are fairly abundant in the crust of the Earth with cerium being the most abundant rare earth and the 25th most-abundant element of all. Overall, elemental abundance in the Earth’s crust are shown in Fig. 1.2, but this does not tell the whole story: the REEs are neither commonly found in isolation nor in very high concentrations. While significant geologic concentrations of iron, nickel, and other elements exist as ores in many places on Earth, the rare earths are rather more uniformly distributed, and where they are locally concentrated, they tend to be found in mixtures and are very hard to separate from each other and also from oxygen. Their chemical similarity is responsible for both their colocation and the difficulty of separating them. Their high affinity for oxygen not only makes it hard to produce metal from their ores but also drives some of their essential uses.

    Fig. 1.2 The abundance of the elements in the earth’s crust, relative to the abundance of silicon. Despite their name the rare earths are not exceptionally scarce. From G.B. Haxel, J.B. Hedrick, G.J. Orris, Rare Earth Elements—Critical Resources for High Technology" (fact sheet), US Geological Survey. https://pubs.usgs.gov/fs/2002/fs087-02/fs087-02.pdf .

    Some of the earliest uses were pioneered by Carl Auer von Welsbach [3]. In many cases former uses have ceased, either because the application became obsolete or alternative materials were identified, but some of Auer’s inventions have had a surprising persistence.

    Lanthanum, cerium, and yttrium oxides, combined with magnesium oxide, formed the first commercial rare earth material, and it was used for gas-lamp mantles. Auer’s 1885 patent on Actinophor was the basis of a business that he started in 1887. Although the material had good emissive power for light and it glowed brightly when heated, the color spectrum was greenish and not considered very attractive, so in 1890 Auer and Haittinger patented an improved gas mantle comprising 99% thoria and 1% ceria [3]. This produced a whiter light, and it became the dominant gas mantle material worldwide, lasting until electric lighting eventually replaced gas: it remains in use today for some gas lanterns such as those used by recreational campers. Thoria is not a rare earth since thorium is an actinide element, not a lanthanide, but the need to find a source for thoria had a profound impact on the production of rare earths.

    Among other uses for the rare earths, carbon arc lamps, especially those used in large movie projectors, were significantly enhanced by additions of lanthanum and other rare earth elements to the carbon electrodes, but like gas lamps, this technology is largely a thing of the past.

    Mischmetal was originally developed by Auer as a mixture of 30% iron and 70% rare earths and today typically contains roughly 50% cerium, 25% lanthanum, and small amounts of neodymium and praseodymium. It is used as a spark generator for cigarette lighters and welding torch igniters—an enduring use over many decades [3] made possible by the highly exothermic reaction of the rare earths with atmospheric oxygen, when a fresh surface is exposed by scraping.

    In the 1920s praseodymium began to be used as a yellowish-orange stain for ceramics and remains in use for this application today. Around the same time, neodymium started to be used to tint glass for both decorative use and for industrial goggles for glassworkers and welders. It also remains in use for this purpose. General Electric’s premium Reveal incandescent lightbulb line used neodymium-tinted glass envelope to provide an improved color balance relative to other incandescent lamps, but this application has been replaced by a new LED version of the Reveal product line, which develops its characteristic spectrum through different means.

    All of these uses of rare earths (with the possible exception of welding goggles) may now be considered to be somewhat inessential, optional, or boutique applications that are either small in volume, are easily replaced, or have otherwise been made obsolete.

    The 1960s, however, saw two new applications for rare earth elements, in which unique benefits emerged from the electronic orbital structure of the lanthanides. These are the first large-scale, initially nonsubstitutable uses for rare earth materials. Zeolite catalysts for crude oil cracking, known as fluid cracking catalysts, or FCCs, are stabilized by the addition of rare earths and achieve significantly longer life as a result of the addition of yttrium and other REEs. The first generation of color televisions suffered from poor color saturation because of the low output of the available red phosphors and the resulting need to mute the green and blue phosphors to maintain a reasonable color balance. This all changed with the discovery of a brighter red phosphor, europium-doped yttrium orthovanadate, which was first adopted by Zenith Electronics, and then industry wide. Truly essential uses for the rare earths were in place for the first time, at least if you believe that color TVs are essential.

    Cerium oxide, or ceria, is used in powder form as a chemical-mechanical planarization medium for silicon wafers and as optician’s rouge for polishing glass, combining excellent abrasive properties with a mild chemical attack of the target materials. It also has widespread applications in catalysis, being used in some automotive catalytic converters and in the coatings of self-cleaning ovens.

    Yttrium oxide, or yttria, has a range of uses including the tough structural ceramic yttria-stabilized zirconia (YSZ) that has many uses including protective coatings on the turbine blades of jet engines, allowing them to run at higher temperatures where they are more efficient. Yttria is the host material for europium and terbium dopants that make red and green phosphors. It is used in dental porcelain and is an essential component in YBa2Cu3O7 high-temperature superconductors. Yttrium is frequently added to metallic alloys to improve their oxidation resistance.

    Other uses of rare earths have come and gone over time. Gadolinium gallium garnet single crystals, for example, were used as substrates for permanent memory chips based on the technology of magnetic bubbles in IBM computers of the 1970s and 1980s.

    Major new uses for rare earth elements emerged in 1967 with the discovery of high-powered permanent magnets based on samarium and cobalt [4] and again in 1982 with the discovery of even stronger magnets based on neodymium, iron, and boron [5, 6]. Today a large fraction of the permanent magnet market depends on the Nd-Fe-B formulation and its derivatives.

    Current uses of the rare earths tend to focus particularly on the properties imparted by their 4f electrons, which are unique to the lanthanides and result in applications related to optical properties, magnetic properties, catalysis, and to some extent mechanical properties. Rare earths are sometimes used to modify the microstructures of major metal products, appearing as grain refiners for castings of aluminum, for example. A partial list of current applications is provided in Table 1.1. In each of these applications, rare earth elements are considered to be difficult or impossible to substitute. Particularly important and persistently unsubstitutable uses for the rare earths include europium, for red light-emitting phosphors, terbium for green light-emitting phosphors, erbium for fiber-optics and medical lasers, and samarium, neodymium, and dysprosium for high-strength permanent magnets. Gadolinium is also important in the control systems of some nuclear reactors, because the uniquely high neutron capture cross section of its nucleus makes it a very effective neutron flux moderator.

    Table 1.1

    Despite the challenges of maintaining a supply chain for rare earth materials, new potential uses continue to emerge, and a partial list of these is provided in Table 1.2. It is possible or even likely, of course, that many of these technologies will fail to penetrate the market as a result of the usual barriers to commercialization (often characterized as the valley of death), but the challenges are increased in cases where an essential material is known to be critical in the sense described in this book, since this tends to deter interest from investors.

    Table 1.2

    Rare earth sources

    As uses for the rare earths emerged around the beginning of the 20th century, sources for them were developed in a variety of places. From 1900 to about 1960, most of the supplies came from monazite ores in placer deposits, associated with alluvial sand. Monazite is nominally either cerium or lanthanum phosphate, but the cation is actually a mixture of different rare earths, and the anion can also be a mixture of phosphate and silicate ions. The mineral can include thorium in addition to the rare earths. Monazite exists as isolated crystals in igneous rocks such as granite, zircon, or ilmenite, and these crystals are released as grains of sand when the rocks weather. The potential of extracting REEs from monazite was identified by Carl Auer von Welsbach, who discovered it in ballast sand in the bilges of a Brazilian ship, as a source of thorium to provide the material for his lamp mantles. While thorium was the initial target of monazite mining, rare earths such as cerium and lanthanum could be extracted, too. Brazil and India dominated the market for monazite until the outbreak of World War II, when South Africa also began production. Australia and the United States (particularly in North Carolina) have also been producers of this ore from time to time.

    New uses for the rare earths that emerged in the years after World War II generated growing demands, and concerns about the radioactivity of thorium created the need to find a different source. A rare earth deposit at Mountain Pass, California, had entered small-scale production in 1952 after being discovered in 1949: this deposit includes both monazite and bastnaesite embedded along with other minerals in a carbonatite matrix. Bastnaesite, bearing rare earths with a much lower thorium content than monazite, became the primary target for rare earth extraction. The Molybdenum Corporation of America, later known as MolyCorp, expanded the production of the bastnaesite ore in the 1960s to meet the growing demand for europium, which was required for the latest consumer electronics sensation—color televisions—and Mountain Pass’s bastnaesite ore body quickly became the world’s dominant source of rare earths.

    Bastnaesite is a fluorocarbonate mineral based on the formula (REE)CO3F, where REE refers to any rare earth element. The Mountain Pass bastnaesite’s rare earth content is dominated by cerium and lanthanum, with decreasing amounts of neodymium, praseodymium, samarium, and other elements with higher atomic number, as shown in Fig. 1.3.

    Fig. 1.3 Relative concentrations of different rare earth elements contained in the Mountain Pass bastnaesite ore.

    In 1977 MolyCorp and its Mountain Pass mine were acquired by Union Oil (Unocal), which became part of Chevron Corporation in 2005. Rare earth demand continued to grow with the development of neodymium-iron-boron permanent magnet alloys in 1982 [5, 6], but the Mountain Pass processing facility was plagued by leaks from its wastewater system, with as many as 60 spills of radioactive waste occurring between 1984 and 1998. In 1984 rare earth production also started at a bastnaesite mine in the Bayan Obo iron mining district near the city of Baotou, Inner Mongolia, in the People’s Republic of China (PRC), and it ramped up very quickly. Production ceased at Mountain Pass in 2002 under the combined pressure of market competition from Baotou and the cost of coping with the challenges of working with hazardous materials in an environmentally sensitive location.

    While bastnaesite provided light rare earths such as lanthanum, cerium, praseodymium, and neodymium, it did not meet the demand for heavy rare earths such as europium, dysprosium, and terbium. These are needed in smaller quantities but are crucial in a range of high-tech applications including efficient lighting, where europium is used to generate red light and terbium emits green light. Dysprosium is added to Nd-Fe-B magnets to improve their performance at elevated temperatures. These heavy rare earths, along with a few others, are mostly obtained from deposits of ion-adsorption clays (also known as lateritic clays) that are fairly widespread in Southern China, although the individual deposits are relatively small. The heavy rare earths are extracted by leaching these clay deposits with acid.

    By 2008 almost 98% of the world’s rare earth supplies came from China. The output of the leading producers at this time is shown in Fig. 1.4.

    Fig. 1.4 The global distribution of rare earth oxide production in 2008. Data from the USGS Mineral Commodity Summary, 2009. https://s3-us-west-2.amazonaws.com/prd-wret/assets/palladium/production/mineral-pubs/mcs/mcs2009.pdf.

    Rare earth supply challenges—2005–15

    In 2005 the Chinese government began imposing annual export quotas on rare earths and reduced the quotas particularly sharply in 2010, as Chinese domestic industries increased their utilization of these materials. In 2005 the export quota represented about 55% of the total Chinese production, and this proportion fell at an accelerating pace until 2010 when the quota was only 23% of the total production, as shown in Fig. 1.5. The actions taken by China were hard to interpret without a clear view of the operation of Chinese commodity markets, so analysts and pundits in the rest of the world ascribed a variety of agendas or goals to China based on the reported actions, leading to concerns about security of supply and a high degree of market nervousness. One impact of the quotas was a marked differential between the prices of rare earths in China and the rest of the world, and the attention of various governments was drawn to the issue as the prices of imported rare earths began to increase and delivery times lengthened. The price history of the rare earths from the mid-2000s to the mid-2010s is illustrated in Fig. 1.6.

    Fig. 1.5 Chinese rare earth oxide production (in green) and export quotas (in red) from 2005 to 2015. The export quotas ended in 2016.

    Fig. 1.6 The prices of three representative rare earths from 2006 to 2016. Source of raw price data: Argus Media Inc. (direct.argusmedia.com).

    In 2008 rare earth industry authorities including Dudley Kingsnorth [7] projected that China’s internal demand for rare earths would outstrip its own production by about 2012, and total world demand would be met only if production outside of China (the so-called rest of the world or ROW) were to increase substantially.

    By this time, concerns about rare earth supplies had already reached high government circles worldwide. Japan is a major importer of rare earths for use in its electronics and electric vehicle industries, and it reacted quickly with government-supported programs to ameliorate potential shortages. Notably a major research and development program to enable recycling of rare metals was started in the summer of 2008, under the banner of Urban Mining [8].

    The European Commission recognized the risks posed to the EU’s economy and issued a communication the EU Parliament in November [9].

    Following various informal consultations, the US Congress held a formal hearing about rare earth supplies, in the Space, Science and Technology Committee of the House of Representatives in March 2010.

    In September 2010 a Chinese fishing vessel collided with a Japanese Coast Guard patrol boat near a group of disputed islands in the East China Sea, known as Senkaku in Japan, and Diaoyu in China. According to China the ship was operating in Chinese waters, but Japan claims the territory, too. The Chinese vessel, its captain, and crew were arrested by the Japanese Coast Guard, resulting in a diplomatic incident between the two nations. While it was widely reported that China cut off rare earth exports to

    Enjoying the preview?
    Page 1 of 1