Main
Lanthanide and Actinide Chemistry (Inorganic Chemistry: A Textbook Series)
Lanthanide and Actinide Chemistry (Inorganic Chemistry: A Textbook Series)
Simon Cotton
Cotton S. / Коттон С.
Henderson W. / Хендерсон У.
Komiya S. / Комия С.
Lawrence G.A. / Лоренц Дж.А.
Muller U. / Мюллер У.
Rehder D. / Редер Д.
Описание:
Список книг: Amouri, H. / Амури, Х. - Chirality in Transition Metal Chemistry. Molecules, Supramolecular Assemblies and Materials (Inorganic Chemistry - A Textbook Series) / Хиральность в химии переходных металлов. Молекулы, супрамолекулярные ансамбли и материалы (Неорганическая химия: серия учебников Wiley)
Cotton S. / Коттон С. - Lanthanide and Actinide Chemistry (Inorganic Chemistry - A Textbook Series) / Химия лантаноидов и актиноидов (Неорганическая химия: серия учебников Wiley)
Henderson W. / Хендерсон У. - Mass Spectrometry of Inorganic, Coordination and Organometallic Compounds - Tools-Techniques-Tips (Inorganic Chemistry - A Textbook series) / Масс-спектрометрия неорганических, координационных и металлорганических соединений (Неорганическая химия: серия учебников Wiley)
Komiya S. / Комия С. - Synthesis of Organometallic Compounds - A Practical Guide (Inorganic Chemistry - A Textbook Series) / Синтез металлорганических соединений (Неорганическая химия: серия учебников Wiley)
Lawrence G.A. / Лоренц Дж.А. - Introduction to Coordination Chemistry (Inorganic Chemistry - A Textbook Series) / Введение в координационную химию (Неорганическая химия: серия учебников Wiley)
Muller U. / Мюллер У. - Inorganic Structural Chemistry (Inorganic Chemistry - A Textbook Series) / Структурная неорганическая химия (Неорганическая химия: серия учебников Wiley)
Rehder D. / Редер Д. - Bioinorganic Vanadium Chemistry (Inorganic Chemistry - A Textbook Series) / Бионеорганическая химия ванадия (Неорганическая химия: серия учебников Wiley)
Henderson W. / Хендерсон У.
Komiya S. / Комия С.
Lawrence G.A. / Лоренц Дж.А.
Muller U. / Мюллер У.
Rehder D. / Редер Д.
Описание:
Список книг: Amouri, H. / Амури, Х. - Chirality in Transition Metal Chemistry. Molecules, Supramolecular Assemblies and Materials (Inorganic Chemistry - A Textbook Series) / Хиральность в химии переходных металлов. Молекулы, супрамолекулярные ансамбли и материалы (Неорганическая химия: серия учебников Wiley)
Cotton S. / Коттон С. - Lanthanide and Actinide Chemistry (Inorganic Chemistry - A Textbook Series) / Химия лантаноидов и актиноидов (Неорганическая химия: серия учебников Wiley)
Henderson W. / Хендерсон У. - Mass Spectrometry of Inorganic, Coordination and Organometallic Compounds - Tools-Techniques-Tips (Inorganic Chemistry - A Textbook series) / Масс-спектрометрия неорганических, координационных и металлорганических соединений (Неорганическая химия: серия учебников Wiley)
Komiya S. / Комия С. - Synthesis of Organometallic Compounds - A Practical Guide (Inorganic Chemistry - A Textbook Series) / Синтез металлорганических соединений (Неорганическая химия: серия учебников Wiley)
Lawrence G.A. / Лоренц Дж.А. - Introduction to Coordination Chemistry (Inorganic Chemistry - A Textbook Series) / Введение в координационную химию (Неорганическая химия: серия учебников Wiley)
Muller U. / Мюллер У. - Inorganic Structural Chemistry (Inorganic Chemistry - A Textbook Series) / Структурная неорганическая химия (Неорганическая химия: серия учебников Wiley)
Rehder D. / Редер Д. - Bioinorganic Vanadium Chemistry (Inorganic Chemistry - A Textbook Series) / Бионеорганическая химия ванадия (Неорганическая химия: серия учебников Wiley)
Categories:
Chemistry\\Inorganic Chemistry
Year:
2006
Edition:
2nd
Language:
english
Pages:
280
ISBN 10:
0470010053
ISBN 13:
9780470010075
Series:
Inorganic Chemistry A Textbook Series
File:
PDF, 2.84 MB
Download (pdf, 2.84 MB)
Preview
- Checking other formats...
- Convert to EPUB
- Convert to FB2
- Convert to MOBI
- Convert to TXT
- Convert to RTF
- Converted file can differ from the original. If possible, download the file in its original format.
- Please login to your account first
-
Need help? Please read our short guide how to send a book to Kindle.The file will be sent to your email address. It may take up to 1-5 minutes before you receive it.The file will be sent to your Kindle account. It may takes up to 1-5 minutes before you received it.
Please note you need to add our NEW email km@bookmail.org to approved e-mail addresses. Read more.
You may be interested in
You can write a book review and share your experiences. Other readers will always be interested in your opinion of the books you've read. Whether you've loved the book or not, if you give your honest and detailed thoughts then people will find new books that are right for them.
1
|
2
|
Lanthanide and Actinide Chemistry Simon Cotton Uppingham School, Uppingham, Rutland, UK Lanthanide and Actinide Chemistry Inorganic Chemistry A Wiley Series of Advanced Textbooks Editorial Board Derek Woollins, University of St. Andrews, UK Bob Crabtree, Yale University, USA David Atwood, University of Kentucky, USA Gerd Meyer, University of Hannover, Germany Previously Published Books In this Series Chemical Bonds: A Dialog Author: J. K. Burdett Bioinorganic Chemistry: Inorganic Elements in the Chemistry of Life – An Introduction and Guide Author: W. Kaim Synthesis of Organometallic Compounds: A Practical Guide Edited by: S. Komiya Main Group Chemistry Second Edition Author: A. G. Massey Inorganic Structural Chemistry Author: U. Muller Stereochemistry of Coordination Compounds Author: A. Von Zelewsky Lanthanide and Actinide Chemistry Author: S. A. Cotton Lanthanide and Actinide Chemistry Simon Cotton Uppingham School, Uppingham, Rutland, UK C 2006 Copyright John Wiley & Sons Ltd, The Atrium, Southern Gate, Chichester, West Sussex PO19 8SQ, England Telephone (+44) 1243 779777 Email (for orders and customer service enquiries): cs-books@wiley.co.uk Visit our Home Page on www.wileyeurope.com or www.wiley.com All Rights Reserved. No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any means, electronic, mechanical, photocopying, recording, scanning or otherwise, except under the terms of the Copyright, Designs and Patents Act 1988 or under the terms of a licence issued by the Copyright Licensing Agency Ltd, 90 Tottenham Court Road, London W1T 4LP, UK, without the permission in writing of the Publisher. Requests to the Publisher should be addressed to the Permissions Department, John Wiley & Sons Ltd, The Atrium, Southern Gate, Chichester, West Sussex PO19 8SQ, England, or emailed to permreq@wiley.co.uk, or faxed to (+44) 1243 770620. Designations used by companies to distinguish their products are often claimed as trademarks. All brand names and product names used in this book are trade names, service marks, trademarks or registered trademarks of their respective owners. The Publisher is not associated with any product or vendor mentioned in this book. This publication is designed to provide accurate and authoritative information in regard to the subject matter covered. It is sold on the understanding that the Publisher is not engaged in rendering professional services. If professional advice or other expert assistance is required, the services of a competent professional should be sought. Other Wiley Editorial Offices John Wiley & Sons Inc., 111 River Street, Hoboken, NJ 07030, USA Jossey-Bass, 989 Market Street, San Francisco, CA 94103-1741, USA Wiley-VCH Verlag GmbH, Boschstr. 12, D-69469 Weinheim, Germany John Wiley & Sons Australia Ltd, 42 McDougall Street, Milton, Queensland 4064, Australia John Wiley & Sons (Asia) Pte Ltd, 2 Clementi Loop #02-01, Jin Xing Distripark, Singapore 129809 John Wiley & Sons Canada Ltd, 22 Worcester Road, Etobicoke, Ontario, Canada M9W 1L1 Wiley also publishes its books in a variety of electronic formats. Some content that appears in print may not be available in electronic books. Library of Congress Cataloging-in-Publication Data Cotton, Simon, Dr. Lanthanide and actinide chemistry / Simon Cotton. p. cm. – (Inorganic chemistry) Includes bibliographical references and index. ISBN-13: 978-0-470-01005-1 (acid-free paper) ISBN-10: 0-470-01005-3 (acid-free paper) ISBN-13: 978-0-470-01006-8 (pbk. : acid-free paper) ISBN-10: 0-470-01006-1 (pbk. : acid-free paper) 1. Rare earth metals. 2. Actinide elements. I. Title. II. Inorganic chemistry (John Wiley & Sons) QD172.R2C68 2006 546 .41—dc22 2005025297 British Library Cataloguing in Publication Data A catalogue record for this book is available from the British Library ISBN-13 9-78-0-470-01005-1 (Cloth) 9-78-0-470-01006-8 (Paper) ISBN-10 0-470-01005-3 (Cloth) 0-470-01006-1 (Paper) Typeset in 10/12pt Times by TechBooks, New Delhi, India Printed and bound in Great Britain by Antony Rowe, Chippenham, Wiltshire This book is printed on acid-free paper responsibly manufactured from sustainable forestry in which at least two trees are planted for each one used for paper production. In memory of Ray and Derek Cotton, my parents. Remember that it was of your parents you were born; how can you repay what they have given to you? (Ecclesiasticus 7.28 RSV) also in memory of Marı́a de los Ángeles Santiago Hernández, a lovely lady and devout Catholic, who died far too young. and to Lisa. Dr Simon Cotton obtained his PhD at Imperial College London. After postdoctoral research and teaching appointments at Queen Mary College, London, and the University of East Anglia, he has taught chemistry in several different schools, and has been at Uppingham School since 1996. From 1984 until 1997, he was Editor of Lanthanide and Actinide Compounds for the Dictionary of Organometallic Compounds and the Dictionary of Inorganic Compounds. He authored the account of Lanthanide Coordination Chemistry for the 2nd edition of Comprehensive Coordination Chemistry (Pergamon) as well as the accounts of Lanthanide Inorganic and Coordination Chemistry for both the 1st and 2nd editions of the Encyclopedia of Inorganic Chemistry (Wiley). His previous books are: S.A. Cotton and F.A. Hart, “The Heavy Transition Elements”, Macmillan, 1975. D.J. Cardin, S.A. Cotton, M. Green and J.A. Labinger, “Organometallic Compounds of the Lanthanides, Actinides and Early Transition Metals”, Chapman and Hall, 1985. S.A. Cotton, “Lanthanides and Actinides”, Macmillan, 1991. S.A. Cotton, “Chemistry of Precious Metals”, Blackie, 1997. Contents Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 Introduction to the Lanthanides . . . . . . . . . . . . 1.1 Introduction . . . . . . . . . . . . . . . . . . . . 1.2 Characteristics of the Lanthanides . . . . . . . . . 1.3 The Occurrence and Abundance of the Lanthanides . 1.4 Lanthanide Ores . . . . . . . . . . . . . . . . . . 1.5 Extracting and Separating the Lanthanides . . . . . 1.5.1 Extraction . . . . . . . . . . . . . . . . . 1.5.2 Separating the Lanthanides . . . . . . . . . 1.6 The Position of the Lanthanides in the Periodic Table 1.7 The Lanthanide Contraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xv . . . . . . . . . . 1 1 2 2 3 3 3 4 6 7 2 The Lanthanides – Principles and Energetics . . . . . . . . . . . . . . . . . . . . . . 2.1 Electron Configurations of the Lanthanides and f Orbitals . . . . . . . . . . . . . . . 2.2 What do f Orbitals Look Like? . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3 How f Orbitals affect Properties of the Lanthanides . . . . . . . . . . . . . . . . . . 2.4 The Lanthanide Contraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5 Electron Configurations of the Lanthanide Elements and of Common Ions . . . . . . . 2.6 Patterns in Ionization Energies . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.7 Atomic and Ionic Radii . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.8 Patterns in Hydration Energies (Enthalpies) for the Lanthanide Ions . . . . . . . . . . 2.9 Enthalpy Changes for the Formation of Simple Lanthanide Compounds . . . . . . . . 2.9.1 Stability of Tetrahalides . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.9.2 Stability of Dihalides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.9.3 Stability of Aqua Ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.10 Patterns in Redox Potentials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 9 10 10 11 12 12 14 14 14 14 17 18 19 3 The Lanthanide Elements and Simple Binary Compounds . . . . . . . . . . . . . . . . 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 The Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.1 Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.2 Synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.3 Alloys and Uses of the Metals . . . . . . . . . . . . . . . . . . . . . . . . . 3.3 Binary Compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.1 Trihalides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.2 Tetrahalides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.3 Dihalides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.4 Oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 23 23 23 24 25 25 25 27 28 29 x Contents Borides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Carbides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Nitrides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hydrides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Sulfides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 31 31 31 32 4 Coordination Chemistry of the Lanthanides . . . . . . . . . . . . . . . . . . . . . . . 4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 Stability of Complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3 Complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3.1 The Aqua Ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3.2 Hydrated Salts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3.3 Other O-Donors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3.4 Complexes of β-Diketonates . . . . . . . . . . . . . . . . . . . . . . . . . 4.3.5 Lewis Base Adducts of β-Diketonate Complexes . . . . . . . . . . . . . . . . 4.3.6 Nitrate Complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3.7 Crown Ether Complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3.8 Complexes of EDTA and Related Ligands . . . . . . . . . . . . . . . . . . . 4.3.9 Complexes of N-Donors . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3.10 Complexes of Porphyrins and Related Systems . . . . . . . . . . . . . . . . . 4.3.11 Halide Complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3.12 Complexes of S-Donors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4 Alkoxides, Alkylamides and Related Substances . . . . . . . . . . . . . . . . . . . . 4.4.1 Alkylamides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4.2 Alkoxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4.3 Thiolates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.5 Coordination Numbers in Lanthanide Complexes . . . . . . . . . . . . . . . . . . . 4.5.1 General Principles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.5.2 Examples of the Coordination Numbers . . . . . . . . . . . . . . . . . . . . 4.5.3 The Lanthanide Contraction and Coordination Numbers . . . . . . . . . . . . 4.5.4 Formulae and Coordination Numbers . . . . . . . . . . . . . . . . . . . . . 4.6 The Coordination Chemistry of the +2 and +4 States . . . . . . . . . . . . . . . . . 4.6.1 The (+2) State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.6.2 The (+4) State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 35 35 37 37 37 38 39 41 41 41 42 43 44 46 46 47 47 48 50 50 50 51 53 54 54 55 56 5 Electronic and Magnetic Properties of the Lanthanides . . . . . . . . . . . . . . . . . 5.1 Magnetic and Spectroscopic Properties of the Ln3+ Ions . . . . . . . . . . . . . . . . 5.2 Magnetic Properties of the Ln3+ Ions . . . . . . . . . . . . . . . . . . . . . . . . . 5.2.1 Adiabatic Demagnetization . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3 Energy Level Diagrams for the Lanthanide Ions, and their Electronic Spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.1 Electronic Spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.2 Hypersensitive Transitions . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4 Luminescence Spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.1 Quenching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.2 Antenna Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.3 Applications of Luminescence to Sensory Probes . . . . . . . . . . . . . . . . 5.4.4 Fluorescence and TV . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.5 Lighting Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 61 62 65 3.4 3.5 3.6 3.7 3.8 65 66 68 69 73 73 74 75 76 Contents xi 5.4.6 Lasers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.7 Euro Banknotes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.5 NMR Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.5.1 β-Diketonates as NMR Shift Reagents . . . . . . . . . . . . . . . . . . . . . 5.5.2 Magnetic Resonance Imaging (MRI) . . . . . . . . . . . . . . . . . . . . . . 5.5.3 What Makes a Good MRI Agent? . . . . . . . . . . . . . . . . . . . . . . . 5.5.4 Texaphyrins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.6 Electron Paramagnetic Resonance Spectroscopy . . . . . . . . . . . . . . . . . . . . 5.7 Lanthanides as Probes in Biological Systems . . . . . . . . . . . . . . . . . . . . . 76 77 77 77 78 79 81 82 83 6 Organometallic Chemistry of the Lanthanides . . . . . . . . . . . . . . . . . . . . . . 6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2 The +3 Oxidation State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2.1 Alkyls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2.2 Aryls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.3 Cyclopentadienyls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.3.1 Compounds of the Unsubstituted Cyclopentadienyl Ligand (C5 H5 = Cp; C5 Me5 = Cp*) . . . . . . . . . . . . . . . . . . . . . . . . . 6.3.2 Compounds [LnCp*3 ] (Cp* = Pentamethylcyclopentadienyl) . . . . . . . . . . 6.3.3 Bis(cyclopentadienyl) Alkyls and Aryls LnCp2 R . . . . . . . . . . . . . . . . 6.3.4 Bis(pentamethylcyclopentadienyl) Alkyls . . . . . . . . . . . . . . . . . . . 6.3.5 Hydride Complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4 Cyclooctatetraene Dianion Complexes . . . . . . . . . . . . . . . . . . . . . . . . 6.5 The +2 State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.5.1 Alkyls and Aryls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.5.2 Cyclopentadienyls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.5.3 Other Compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.6 The +4 State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.7 Metal–Arene Complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.8 Carbonyls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 89 89 89 90 91 91 94 95 96 98 98 99 99 100 100 101 101 102 7 The Misfits: Scandium, Yttrium, and Promethium . . . . . . . . . . . . . . . . . . . . 7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.2 Scandium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.2.1 Binary Compounds of Scandium . . . . . . . . . . . . . . . . . . . . . . . 7.3 Coordination Compounds of Scandium . . . . . . . . . . . . . . . . . . . . . . . . 7.3.1 The Aqua Ion and Hydrated Salts . . . . . . . . . . . . . . . . . . . . . . . 7.3.2 Other Complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.3.3 Alkoxides and Alkylamides . . . . . . . . . . . . . . . . . . . . . . . . . . 7.3.4 Patterns in Coordination Number . . . . . . . . . . . . . . . . . . . . . . . 7.4 Organometallic Compounds of Scandium . . . . . . . . . . . . . . . . . . . . . . . 7.5 Yttrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.6 Promethium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 107 107 108 108 108 109 111 112 114 114 115 8 The Lanthanides and Scandium in Organic Chemistry . . . . . . . . . . . . . . . . . . 8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.2 Cerium(IV) Compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.2.1 Oxidation of Aromatics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.2.2 Oxidation of Alkenes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.2.3 Oxidation of Alcohols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 121 121 121 123 123 xii Contents 8.3 Samarium(II) Iodide, SmI2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.1 Reduction of Halogen Compounds . . . . . . . . . . . . . . . . . . . . . . . 8.3.2 Reduction of α-Heterosubstituted Ketones . . . . . . . . . . . . . . . . . . . 8.3.3 Reductions of Carbonyl Groups . . . . . . . . . . . . . . . . . . . . . . . . 8.3.4 Barbier Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.5 Reformatsky Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.4 Lanthanide β-Diketonates as Diels–Alder Catalysts . . . . . . . . . . . . . . . . . . 8.4.1 Hetero-Diels–Alder Reactions . . . . . . . . . . . . . . . . . . . . . . . . . 8.5 Cerium(III) Chloride and Organocerium Compounds . . . . . . . . . . . . . . . . . 8.6 Cerium(III) Chloride and Metal Hydrides . . . . . . . . . . . . . . . . . . . . . . . 8.7 Scandium Triflate and Lanthanide Triflates . . . . . . . . . . . . . . . . . . . . . . 8.7.1 Friedel–Crafts Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.7.2 Diels–Alder Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.7.3 Mannich Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.7.4 Imino-Diels–Alder Reactions . . . . . . . . . . . . . . . . . . . . . . . . . 8.8 Alkoxides and Aryloxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.9 Lanthanide Metals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.10 Organometallics and Catalysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124 124 124 124 126 127 127 128 128 130 131 131 132 132 133 134 135 136 9 Introduction to the Actinides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.1 Introduction and Occurrence of the Actinides . . . . . . . . . . . . . . . . . . . . . 9.2 Synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3 Extraction of Th, Pa, and U . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.1 Extraction of Thorium . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.2 Extraction of Protactinium . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.3 Extraction and Purification of Uranium . . . . . . . . . . . . . . . . . . . . 9.4 Uranium Isotope Separation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.4.1 Gaseous Diffusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.4.2 Gas Centrifuge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.4.3 Electromagnetic Separation . . . . . . . . . . . . . . . . . . . . . . . . . . 9.4.4 Laser Separation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.5 Characteristics of the Actinides . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.6 Reduction Potentials of the Actinides . . . . . . . . . . . . . . . . . . . . . . . . . 9.7 Relativistic Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145 145 145 147 147 148 148 148 148 148 149 149 149 150 152 10 Binary Compounds of the Actinides . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.2 Halides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.2.1 Syntheses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.2.2 Structure Types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.3 Thorium Halides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.4 Uranium Halides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.4.1 Uranium(VI) Compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.4.2 Uranium(V) Compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.4.3 Uranium (IV) Compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.4.4 Uranium(III) Compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.4.5 Uranium Hexafluoride and Isotope Separation . . . . . . . . . . . . . . . . . 155 155 155 156 158 159 160 161 161 161 163 163 Contents xiii The Actinide Halides (Ac–Am) excluding U and Th . . . . . . . . . . . . . . . . . 10.5.1 Actinium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.5.2 Protactinium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.5.3 Neptunium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.5.4 Plutonium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.5.5 Americium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Halides of the Heavier Transactinides . . . . . . . . . . . . . . . . . . . . . . . . 10.6.1 Curium(III) Chloride . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.6.2 Californium(III) Chloride, Californium(III) Iodide, and Californium(II) Iodide . 10.6.3 Einsteinium(III) Chloride . . . . . . . . . . . . . . . . . . . . . . . . . . Oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.7.1 Thorium Oxide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.7.2 Uranium Oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Uranium Hydride UH3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Oxyhalides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165 165 165 166 166 167 168 168 168 168 169 169 169 170 170 11 Coordination Chemistry of the Actinides . . . . . . . . . . . . . . . . . . . . . . . . . 11.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.2 General Patterns in the Coordination Chemistry of the Actinides . . . . . . . . . . . 11.3 Coordination Numbers in Actinide Complexes . . . . . . . . . . . . . . . . . . . . 11.4 Types of Complex Formed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.5 Uranium and Thorium Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . 11.5.1 Uranyl Complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.5.2 Coordination Numbers and Geometries in Uranyl Complexes . . . . . . . . 11.5.3 Some Other Complexes . . . . . . . . . . . . . . . . . . . . . . . . . . 11.5.4 Uranyl Nitrate and its Complexes; their Role in Processing Nuclear Waste . . 11.5.5 Nuclear Waste Processing . . . . . . . . . . . . . . . . . . . . . . . . . 11.6 Complexes of the Actinide(IV) Nitrates and Halides . . . . . . . . . . . . . . . . . 11.6.1 Thorium Nitrate Complexes . . . . . . . . . . . . . . . . . . . . . . . . 11.6.2 Uranium(IV) Nitrate Complexes . . . . . . . . . . . . . . . . . . . . . . 11.6.3 Complexes of the Actinide(IV) Halides . . . . . . . . . . . . . . . . . . . 11.7 Thiocyanates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.8 Amides, Alkoxides and Thiolates . . . . . . . . . . . . . . . . . . . . . . . . . . 11.8.1 Amide Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.8.2 Alkoxides and Aryloxides . . . . . . . . . . . . . . . . . . . . . . . . . 11.9 Chemistry of Actinium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.10 Chemistry of Protactinium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.11 Chemistry of Neptunium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.11.1 Complexes of Neptunium . . . . . . . . . . . . . . . . . . . . . . . . . 11.12 Chemistry of Plutonium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.12.1 Aqueous Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.12.2 The Stability of the Oxidation States of Plutonium . . . . . . . . . . . . . . 11.12.3 Coordination Chemistry of Plutonium . . . . . . . . . . . . . . . . . . . 11.12.4 Plutonium in the Environment . . . . . . . . . . . . . . . . . . . . . . . 11.13 Chemistry of Americium and Subsequent Actinides . . . . . . . . . . . . . . . . . 11.13.1 Potentials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.14 Chemistry of the Later Actinides . . . . . . . . . . . . . . . . . . . . . . . . . . 173 173 173 174 174 175 175 177 178 179 179 180 180 180 181 183 184 184 185 186 187 188 188 189 189 190 191 193 195 195 196 10.5 10.6 10.7 10.8 10.9 xiv Contents 12 Electronic and Magnetic Properties of the Actinides . . . . . . . . . . . . . . . . . . . 12.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.2 Absorption Spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.2.1 Uranium(VI) – UO2 2+ – f 0 . . . . . . . . . . . . . . . . . . . . . . . . . 12.2.2 Uranium(V) – f 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.2.3 Uranium(IV) – f 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.2.4 Spectra of the Later Actinides . . . . . . . . . . . . . . . . . . . . . . . 12.3 Magnetic Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201 201 202 202 203 204 206 207 13 Organometallic Chemistry of the Actinides . . . . . . . . . . . . . . . . . . . . . . . 13.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.2 Simple σ-Bonded Organometallics . . . . . . . . . . . . . . . . . . . . . . . . . 13.3 Cyclopentadienyls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.3.1 Oxidation State (VI) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.3.2 Oxidation State (V) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.3.3 Oxidation State (IV) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.3.4 Oxidation State (III) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.4 Compounds of the Pentamethylcyclopentadienyl Ligand (C5 Me5 = Cp*) . . . . . . . 13.4.1 Oxidation State(IV) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.4.2 Cationic Species and Catalysts . . . . . . . . . . . . . . . . . . . . . . . 13.4.3 Hydrides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.4.4 Oxidation State (III) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.5 Tris(pentamethylcyclopentadienyl) Systems . . . . . . . . . . . . . . . . . . . . . 13.6 Other Metallacycles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.7 Cyclooctatetraene Dianion Compounds . . . . . . . . . . . . . . . . . . . . . . . 13.8 Arene Complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.9 Carbonyls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209 209 209 211 211 211 211 214 215 215 216 217 218 219 219 219 221 222 14 Synthesis of the Transactinides and their Chemistry . . . . . . . . . . . . . . . . . . . 14.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.2 Finding New Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.3 Synthesis of the Transactinides . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.4 Naming the Transactinides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.5 Predicting Electronic Arrangements . . . . . . . . . . . . . . . . . . . . . . . . . 14.6 Identifying the Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.7 Predicting Chemistry of the Transactinides . . . . . . . . . . . . . . . . . . . . . 14.8 What is known about the Chemistry of the Transactinides . . . . . . . . . . . . . . 14.8.1 Element 104 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.8.2 Element 105 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.8.3 Element 106 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.8.4 Element 107 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.8.5 Element 108 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.9 And the Future? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225 225 226 226 229 230 230 233 234 234 234 234 235 235 236 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237 Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253 Preface This book is aimed at providing a sound introduction to the chemistry of the lanthanides, actinides and transactinides to undergraduate students. I hope that it will also be of value to teachers of these courses. Whilst not being anything resembling a comprehensive monograph, it does attempt to give a factual basis to the area, and the reader can use a fairly comprehensive bibliography to range further. Since I wrote a previous book in this area (1991), the reader may wonder why on earth I have bothered again. The world of f-block chemistry has moved on. It is one of active and important research, with names like Bünzli, Evans, Ephritikhine, Lappert, Marks and Parker familiar world-wide (I am conscious of names omitted). Not only have several more elements been synthesized (and claims made for others), but lanthanides and their compounds are routinely employed in many areas of synthetic organic chemistry; gadolinium compounds find routine application in MRI scans; and there are other spectroscopic applications, notably in luminescence. Whilst some areas are hardly changed, at this level at least (e.g. actinide magnetism and spectroscopy), a lot more compounds have been described, accounting for the length of the chapters on coordination and organometallic chemistry. I have tried to spell out the energetics of lanthanide chemistry in more detail, whilst I have provided some end-of-chapter questions, of variable difficulty, which may prove useful for tutorials. I have supplied most, but not all, of the answers to these (my answers, which are not always definitive). It is a pleasure to thank all those who have contributed to the book: Professor Derek Woollins, for much encouragement at different stages of the project; Professor James Anderson, for many valuable comments on Chapter 8; Martyn Berry, who supplied valuable comment on early versions of several chapters; to Professors Michel Ephritikhine, Allan White and Jack Harrowfield, and Dr J.A.G. Williams, and many others, for exchanging e-mails, correspondence and ideas. I’m very grateful to Dr Mary P. Neu for much information on plutonium. The staff of the Libraries of the Chemistry Department of Cambridge University and of the Royal Society of Chemistry, as well as the British Library, have been quite indispensable in helping with access to the primary literature. I would also wish to thank a number of friends – once again Dr Alan Hart, who got me interested in lanthanides in the first place; Professor James Anderson (again), Dr Andrew Platt, Dr John Fawcett, and Professor Paul Raithby, for continued research collaboration and obtaining spectra and structures from unpromising crystals, so that I have kept a toe-hold in the area. Over the last 8 years, a number of Uppingham 6th form students have contributed to my efforts in lanthanide coordination chemistry – John Bower, Oliver Noy, Rachel How, Vilius Franckevicius, Leon Catallo, Franz Niecknig, Victoria Fisher, Alex Tait and Joanna Harris. Finally, thanks are most certainly due to Dom Paul-Emmanuel Clénet and the Benedictine community of the Abbey of Bec, for continued hospitality during several Augusts when I have been compiling the book. Simon Cotton 1 Introduction to the Lanthanides By the end of this chapter you should be able to: r understand that lanthanides differ in their properties from the s- and d-block metals; r recall characteristic properties of these elements; r appreciate reasons for their positioning in the Periodic Table; r understand how the size of the lanthanide ions affects certain properties and how this can be used in the extraction and separation of the elements; r understand how to obtain pure samples of individual Ln3+ ions. 1.1 Introduction Lanthanide chemistry started in Scandinavia. In 1794 Johann Gadolin succeeded in obtaining an ‘earth’(oxide) from a black mineral subsequently known as gadolinite; he called the earth yttria. Soon afterwards, M.H. Klaproth, J.J. Berzelius and W. Hisinger obtained ceria, another earth, from cerite. However, it was not until 1839–1843 that the Swede C.G. Mosander first separated these earths into their component oxides; thus ceria was resolved into the oxides of cerium and lanthanum and a mixed oxide ‘didymia’ (a mixture of the oxides of the metals from Pr through Gd). The original yttria was similarly separated into substances called erbia, terbia, and yttria (though some 40 years later, the first two names were to be reversed!). This kind of confusion was made worse by the fact that the newly discovered means of spectroscopic analysis permitted misidentifications, so that around 70 ‘new’ elements were erroneously claimed in the course of the century. Nor was Mendeleev’s revolutionary Periodic Table a help. When he first published his Periodic Table in 1869, he was able to include only lanthanum, cerium, didymium (now known to have been a mixture of Pr and Nd), another mixture in the form of erbia, and yttrium; unreliable information about atomic mass made correct positioning of these elements in the table difficult. Some had not yet been isolated as elements. There was no way of predicting how many of these elements there would be until Henry Moseley (1887–1915) analysed the X-ray spectra of elements and gave meaning to the concept of atomic number. He showed that there were 15 elements from lanthanum to lutetium (which had only been identified in 1907). The discovery of radioactive promethium had to wait until after World War 2. It was the pronounced similarity of the lanthanides to each other, especially each to its neighbours (a consequence of their general adoption of the +3 oxidation state in aqueous solution), that caused their classification and eventual separation to be an extremely difficult undertaking. Lanthanide and Actinide Chemistry S. Cotton C 2006 John Wiley & Sons, Ltd. 2 Introduction to the Lanthanides Subsequently it was not until the work of Bohr and of Moseley that it was known precisely how many of these elements there were. Most current versions of the Periodic Table place lanthanum under scandium and yttrium. 1.2 Characteristics of the Lanthanides The lanthanides exhibit a number of features in their chemistry that differentiate them from the d-block metals. The reactivity of the elements is greater than that of the transition metals, akin to the Group II metals: 1. A very wide range of coordination numbers (generally 6–12, but numbers of 2, 3 or 4 are known). 2. Coordination geometries are determined by ligand steric factors rather than crystal field effects. 3. They form labile ‘ionic’ complexes that undergo facile exchange of ligand. 4. The 4f orbitals in the Ln3+ ion do not participate directly in bonding, being well shielded by the 5s2 and 5p6 orbitals. Their spectroscopic and magnetic properties are thus largely uninfluenced by the ligand. 5. Small crystal-field splittings and very sharp electronic spectra in comparison with the d-block metals. 6. They prefer anionic ligands with donor atoms of rather high electronegativity (e.g. O, F). 7. They readily form hydrated complexes (on account of the high hydration energy of the small Ln3+ ion) and this can cause uncertainty in assigning coordination numbers. 8. Insoluble hydroxides precipitate at neutral pH unless complexing agents are present. 9. The chemistry is largely that of one (3+) oxidation state (certainly in aqueous solution). 10. They do not form Ln=O or Ln≡N multiple bonds of the type known for many transition metals and certain actinides. 11. Unlike the transition metals, they do not form stable carbonyls and have (virtually) no chemistry in the 0 oxidation state. 1.3 The Occurrence and Abundance of the Lanthanides Table 1.1 presents the abundance of the lanthanides in the earth’s crust and in the solar system as a whole. (Although not in the same units, the values in each list are internally consistent.) Two patterns emerge from these data. First, that the lighter lanthanides are more abundant than the heavier ones; secondly, that the elements with even atomic number are more abundant than those with odd atomic number. Overall, cerium, the most abundant lanthanide Table 1.1 Abundance of the lanthanides Crust (ppm) Solar System (with respect to 107 atoms Si) La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu Y 35 4.5 66 1.2 9.1 1.7 40 8.5 0.0 0.0 7 2.5 2.1 1.0 6.1 3.3 1.2 0.6 4.5 3.9 1.3 0.9 3.5 2.5 0.5 0.4 3.1 2.4 0.8 0.4 31 40.0 Extracting and Separating the Lanthanides 3 on earth, has a similar crustal concentration to the lighter Ni and Cu, whilst even Tm and Lu, the rarest lanthanides, are more abundant than Bi, Ag or the platinum metals. The abundances are a consequence of how the elements were synthesized by atomic fusion in the cores of stars with heavy elements only made in supernovae. Synthesis of heavier nuclei requires higher temperature and pressures and so gets progressively harder as the atomic number increases. The odd/even alternation (often referred to as the Oddo–Harkins rule) is again general, and reflects the facts that elements with odd mass numbers have larger nuclear capture cross sections and are more likely to take up another neutron, so elements with odd atomic number (and hence odd mass number) are less common than those with even mass number. Even-atomic-number nuclei are more stable when formed. 1.4 Lanthanide Ores Principal sources (Table 1.2) are the following: Bastnasite LnFCO3 ; Monazite (Ln, Th)PO4 (richer in earlier lanthanides); Xenotime (Y, Ln)PO4 (richer in later lanthanides). In addition to these, there are Chinese rare earth reserves which amount to over 70% of the known world total, mainly in the form of the ionic ores from southern provinces. These Chinese ion-absorption ores, weathered granites with lanthanides adsorbed onto the surface of aluminium silicates, are in some cases low in cerium and rich in the heavier lanthanides (Longnan) whilst the Xunwu deposits are rich in the lighter metals; the small particle size makes them easy to mine. The Chinese ores have made them a leading player in lanthanide chemistry. Table 1.2 Typical abundance of the lanthanides in oresa % La Monazite Bastnasite Xenotime 20 43 4.5 16 0 33.2 49.1 4.3 12 0 0.5 5 0.7 2.2 0 a Bold Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu Y 3 0.1 1.5 0.05 0.6 0.05 0.2 0.02 0.1 0.02 2.5 0.8 0.12 0.17 160 310 50 35 8 6 1 0.1 1.9 0.2 4 1 8.6 2 5.4 0.9 6.2 0.4 60.0 values are in ppm. 1.5 Extracting and Separating the Lanthanides These two processes are not necessarily coterminous. Whilst electronic, optical and magnetic applications require individual pure lanthanides, the greatest quantity of lanthanides is used as mixtures, e.g. in mischmetal or oxide catalysts. 1.5.1 Extraction After initial concentration by crushing, grinding and froth flotation, bastnasite is treated with 10% HCl to remove calcite, by which time the mixture contains around 70% lanthanide oxides. This is roasted to oxidize the cerium content to CeIV ; on further extraction with HCl, the Ce remains as CeO2 , whilst the lanthanides in the (+3) state dissolve as a solution of the chlorides. Monazite is usually treated with NaOH at 150 ◦ C to remove phosphate as Na3 PO4 , leaving a mixture of the hydrated oxides, which are dissolved in boiling HCl at pH 3.5, separating the lanthanides from insoluble ThO2 . Sulfuric acid can also be used to dissolve the lanthanides. 4 Introduction to the Lanthanides 1.5.2 Separating the Lanthanides These can be divided into four types: chemical separations, fractional crystallization, ionexchange methods and solvent extraction. Of these, only the last-named is used on a commercial scale (apart from initial separation of cerium). Chemical separations rely on using stabilities of unusual oxidation states; thus Eu2+ is the only ion in that oxidation state formed on reduction by zinc amalgam and can then be precipitated as EuSO4 (note the similarity with heavier Group 2 metals). Repeated (and tedious) fractional crystallization, which made use of slight solubility differences between the salts of neighbouring lanthanides, such as the bromates Ln(BrO3 )3 .9H2 O, ethyl sulfates and double nitrates, were once the only possible way of obtaining pure lanthanides, as with the 15 000 recrystallizations carried out by the American C. James to get pure thulium bromate (1911) (Figure 1.1 indicates the principle of this method). original Legend solution = dissolve in H2O = evaporate to crystals = mother liquor A1 = crystals = combine B1 A2 D1 ble olu t s uent as Le nstit co E1 B2 C2 D2 E2 A3 B3 C3 D3 A4 B4 C4 A5 B5 Mo co st so ns lu titu ble en t C1 Middle fraction Figure 1.1 Diagrammatic representation of the system of fractional crystallization used to separate salts of the rare-earth elements (reproduced with permission from D.M. Yost, H. Russell and C.S. Garner, The Rare Earth Elements and their Compounds, John Wiley, 1947.) Ion-exchange chromatography is not of real commercial importance for large-scale production, but historically it was the method by which fast high-purity separation of the lanthanides first became feasible. As radioactive lanthanide isotopes are important fission products of the fission of 235 U and therefore need to be separated from uranium, and because the actinides after plutonium tend to resemble the lanthanides, the development of the technique followed on the Manhattan project. It was found that if Ln3+ ions were adsorbed at the top of a cation-exchange resin, then treated with a complexing agent such as buffered citric acid, then the cations tended to be eluted in reverse atomic number order (Figure 1.2a); the anionic ligand binds most strongly to the heaviest (and smallest) cation, which has the highest charge density. A disadvantage of this approach when scaled up to high concentration is that the peaks tend to overlap (Figure 1.2b). It was subsequently found that amine polycarboxylates such as EDTA4− gave stronger complexes and much better separations. In practice, some Cu2+ ions (‘retainer’) are added to prevent precipitation of either the free acid H4 EDTA or the lanthanide complex Extracting and Separating the Lanthanides 105 5 Ho Lu Concentration – mg. oxide/ liter Concentration – counts/ min. Tb Y Tm Dy 101 0 2500 150 Sm Nd Pr 0 0 Elution Time – min. (a) 60 Volume of Eluate – liters (b) Figure 1.2 (a) Cation-exchange chromatography of lanthanides, (b) overlap of peaks at high concentration. (a) Tracer-scale elution with 5% citrate at pH 3.20 (redrawn from B.H. Ketelle and G.E. Boyd, J. Am. Chem. Soc., 1947, 69, 2800). (b) Macro-scale elution with 0.1% citrate at pH 5.30 (redrawn from F.H. Spedding, E.I. Fulmer, J.E. Powell, and T.A. Butler, J. Am. Chem. c Soc., 1950, 72, 2354). Reprinted with permission of the American Chemical Society 1978. HLn(EDTA).xH2 O on the resin. The major disadvantage of this method is that it is a slow process for large-scale separations. Solvent extraction has come to be used for the initial stage of the separation process, to give material with up to 99.9% purity. In 1949, it was found that Ce4+ could readily be separated from Ln3+ ions by extraction from a solution in nitric acid into tributyl phosphate [(BuO)3 PO]. Subsequently the process was extended to separating the lanthanides, using a non-polar organic solvent such as kerosene and an extractant such as (BuO)3 PO or bis (2-ethylhexyl)phosphinic acid [[C4 H9 CH(C2 H5 )]2 P=O(OH)] to extract the lanthanides from aqueous nitrate solutions. The heavier lanthanides form complexes which are more soluble in the aqueous layer. After the two immiscible solvents have been agitated together and separated, the organic layer is treated with acid and the lanthanide extracted. The solvent is recycled and the aqueous layer put through further stages. For a lanthanide LnA distributed between two phases, a distribution coefficent DA is defined: DA = [LnA in organic phase] / [LnA in aqueous phase] For two lanthanides LnA and LnB in a mixture being separated, a separation factor βBA can be defined, where βBA = DA /DB β is very close to unity for two adjacent lanthanides in the Periodic Table (obviously, the larger β is, the better the separation). In practice this process is run using an automated continuous counter-current circuit in which the organic solvent flows in the opposite direction to the aqueous layer containing the lanthanides. An equilibrium is set up between the lanthanide ions in the aqueous phase and the organic layer, with there tending to be a relative enhancement of the concentration of the heavier lanthanides in the organic layer. Because the separation between adjacent 6 Introduction to the Lanthanides Back-extract (water) Scrub 5M HNO3 Back extraction 10 stages Stripped solvent for re-cycling Scrub Solvent containing Sm-Nd-HNO3 Aqueous solution Pr-La Precipitate and re-dissolve Concentrate or precipitate and re-dissolve Solvent Feed Stripped solvent Aqueous Sm, etc. Scrub Solvent Feed Back-extract Scrub and Extraction Solvent Sm Solvent (Undiluted tributyl phosphate) Aqueous solution Sm-Nd Back extract Back extraction Feed Saturated neutral nitrate solution of La-Pr-Nd-Sm Scrub and extraction Perhaps 20 stages Back extraction Scrub and extraction Solvent Pr Stripped solvent Aqueous Aqueous Pr Nd Aqueous La Figure 1.3 Schematic diagram of lanthanon separation by solvent extraction. From R.J. Callow, The Rare Earth Industry, Pergamon, 1966; reproduced by permission. lanthanides in each exchange is relatively slight, over a thousand exchanges are used (see Figure 1.3). This method affords lanthanides of purity up to the 99.9% purity level and is thus well suited to large-scale separation, the products being suited to ordinary chemical use. However, for electronic or spectroscopic use (‘phosphor grade’) 99.999% purity is necessary, and currently ion-exchange is used for final purification to these levels. The desired lanthanides are precipitated as the oxalate or hydroxide and converted into the oxides (the standard starting material for many syntheses) by thermal decomposition. Various other separation methods have been described, one recent one involving the use of supercritical carbon dioxide at 40 ◦ C and 100 atm to convert the lanthanides into their carbonates whilst the quadrivalent metals (e.g. Th and Ce) remain as their oxides. 1.6 The Position of the Lanthanides in the Periodic Table As already mentioned, neither Mendeleev nor his successors could ‘place’ the lanthanides in the Periodic Table. Not only was there no recognizable atomic theory until many years afterwards, but, more relevant to how groupings of elements were made in those days, there was no comparable block of elements for making comparisons. The lanthanides were sui generis. The problem was solved by the combined (but separate) efforts of Moseley and Bohr, the former showing that La–Lu was composed of 15 elements with atomic numbers from 57 to 71, whilst the latter concluded that the fourth quantum shell could accommodate 32 electrons, and that the lanthanides were associated with placing electrons into the 4f orbitals. The Lanthanide Contraction 7 The Periodic Table places elements in atomic number order, with the lanthanides falling between barium (56) and hafnium (72). For reasons of space, most present-day Periodic Tables are presented with Groups IIA and IVB (2 and 4) separated only by the Group IIIB (3) elements. Normally La (and Ac) are grouped with Sc and Y, but arguments have been advanced for an alternative format, in which Lu (and Lr) are grouped with Sc and Y (see e.g. W.B. Jensen, J. Chem. Educ., 1982, 59, 634) on the grounds that trends in properties (e.g. atomic radius, I.E., melting point) in the block Sc-Y-Lu parallel those in the Group Ti-Zr-Hf rather closely, and that there are resemblances in the structures of certain binary compounds. Certainly on size grounds, Lu resembles Y and Sc (it is intermediate in size between them) rather more than does La, owing to the effects of the ‘lanthanide contraction’. The resemblances between Sc and Lu are, however, by no means complete. 1.7 The Lanthanide Contraction The basic concept is that there is a decrease in radius of the lanthanide ion Ln3+ on crossing the series from La to Lu. This is caused by the poor screening of the 4f electrons. This causes neighbouring lanthanides to have similar, but not identical, properties, and is discussed in more detail in Section 2.4. Question 1.1 Using the information you have been given in Section 1.2, draw up a table comparing (in three columns) the characteristic features of the s-block metals (use group 1 as typical) and the d-block transition metals. Answer 1.1 see Table 1.3 for one such comparison. Table 1.3 Comparison of 4f, 3d and Group I metals Electron configurations of ions Stable oxidation states Coordination numbers in complexes Coordination polyhedra in complexes Trends in coordination numbers Donor atoms in complexes Hydration energy Ligand exchange reactions Magnetic properties of ions Electronic spectra of ions Crystal field effects in complexes Organometallic compounds Organometallics in low oxidation states Multiply bonded atoms in complexes 4f 3d Group I Variable Variable Noble gas Usually +3 Commonly 8–10 Variable Usually 6 1 Often 4–6 Minimise repulsion Directional Minimise repulsion Often constant in block Often constant in block Increase down group ‘Hard’ preferred High Usually fast Independent of environment Sharp lines Weak ‘Hard’ and ‘soft’ Usually moderate Fast and slow Depends on environment and ligand field Broad lines Strong ‘Hard’ preferred Low Fast None None None Usually ionic, some with covalent character Few Covalently bonded Ionically bonded Common None None Common None 2 The Lanthanides – Principles and Energetics By the end of this chapter you should be able to: r recognise the difference between f-orbitals and other types of orbitals; r understand that they are responsible for the particular properties of the lanthanides; r give the electron configurations of the lanthanide elements and Ln3+ ions; r explain the reason for the lanthanide contraction; r understand the effect of the lanthanide contraction upon properties of the lanthanides and subsequent elements; r explain patterns in properties such as ionization and hydration energies; r recall that lanthanides behave similarly when there is no change in the 4f electron popr r ulation, but that they differ when the change involves a change in the number of 4f electrons; relate the stability of oxidation states to the ionization energies; calculate enthalpy changes for the formation of the aqua ions and of the lanthanide halides and relate these to the stability of particular compounds. 2.1 Electron Configurations of the Lanthanides and f Orbitals The lanthanides (and actinides) are those in which the 4f (and 5f) orbitals are gradually filled. At lanthanum, the 5d subshell is lower in energy than 4f, so lanthanum has the electron configuration [Xe] 6s2 5d1 (Table 2.1). As more protons are added to the nucleus, the 4f orbitals contract rapidly and become more stable than the 5d (as the 4f orbitals penetrate the ‘xenon core’ more) (see Figure 2.1), so that Ce has the electron configuration [Xe] 6s2 5d1 4f1 and the trend continues with Pr having the arrangement [Xe] 6 s2 4f 3 . This pattern continues for the metals Nd–Eu, all of which have configurations [Xe] 6s2 4f n (n = 4–7) After europium, the stability of the halffilled f subshell is such that the next electron is added to the 5d orbital, Gd being [Xe] 6s2 5d1 4f 7 ; at terbium, however, the earlier pattern is resumed, with Tb having the configuration [Xe] 6s2 4f 9 , and succeeding elements to ytterbium being [Xe] 6s2 4f n (n = 10–14). The last lanthanide, lutetium, where the 4f subshell is now filled, is predictably [Xe] 6s2 5d1 4f 14 . Lanthanide and Actinide Chemistry S. Cotton C 2006 John Wiley & Sons, Ltd. 10 The Lanthanides – Principles and Energetics Table 2.1 Electron configurations of the lanthanides and their common ions La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu Y Atom Ln3+ [Xe] 5d1 6s2 [Xe] 4f1 5d1 6s2 [Xe] 4f3 6s2 [Xe] 4f4 6s2 [Xe] 4f5 6s2 [Xe] 4f6 6s2 [Xe] 4f7 6s2 [Xe] 4f7 5d1 6s2 [Xe] 4f9 6s2 [Xe] 4f10 6s2 [Xe] 4f11 6s2 [Xe] 4f12 6s2 [Xe] 4f13 6s2 [Xe] 4f14 6s2 [Xe] 4f14 5d1 6s2 [Kr] 4d1 5s2 [Xe] [Xe] 4f1 [Xe] 4f2 [Xe] 4f3 [Xe] 4f4 [Xe] 4f5 [Xe] 4f6 [Xe] 4f7 [Xe] 4f8 [Xe] 4f9 [Xe] 4f10 [Xe] 4f11 [Xe] 4f12 [Xe] 4f13 [Xe] 4f14 [Kr] Ln4+ Ln2+ [Xe] [Xe] 4f1 [Xe] 4f2 [Xe] 4f4 [Xe] 4f6 [Xe] 4f7 [Xe] 4f7 [Xe] 4f8 [Xe] 4f10 [Xe] 4f13 [Xe] 4f14 Probability 4f 5d 6s r Figure 2.1 The radial part of the hydrogenic wave functions for the 4f, 5d and 6s orbitals of cerium (after H.G. Friedman et al. J. Chem. Educ. 1964, 41, 357). Reproduced by permission of the American Chemical c 1964. Society 2.2 What do f Orbitals Look Like? They are generally represented in one of two ways, either as a cubic set, or as a general set, depending upon which way the orbitals are combined. The cubic set comprises fxyz ; fz(x2−y2) , fz(y2−z2) and fy(z2−x2); fz3 , fx3 and fy3 . The general set, more useful in non-cubic environments, uses a different combination: fz3 ; fxz2 and fyz2 ; fxyz ; fz(x2−y2) , fx(x2−3y2) and f y(3x2−y2) ; Figure 2.2 shows the general set. 2.3 How f Orbitals affect Properties of the Lanthanides The 4f orbitals penetrate the xenon core appreciably. Because of this, they cannot overlap with ligand orbitals and therefore do not participate significantly in bonding. As a result of The Lanthanide Contraction 11 z z + − − + + − x (a) fz 3 (b) fx z 2 y z − − + + + − − + + x x x − y (c) fx (x 2 − 3y 2) (d) fxyz Figure 2.2 (a) fz 3 , (fx 3 and f y 3 are similar, extending along the x- and y-axes repectively); (b) fx z 3 , (f yz 3 is similar, produced by a 90◦ rotation about the z-axis); (c) fx(x 2 −3y 2 ) , f y(3x 2 −y 2 ) is similar, formed by a 90◦ clockwise rotation round the z-axis); (d) fx yz , (fx z 2 −y 2 ), f y(z 2 −y 2 ) and fz(x 2 −y 2 ) are produced by a 45◦ rotation about the x, y and z-axes respectively). The cubic set comprises fx 3 , f y 3 , fz 3 , fx yz , fx(z 2 −y 2 ) f y(z 2 −x 2 ) and fz(x 2 −y 2 ) ; the general set is made of fz 3 , fx z 2 , f yz 2 , fx yz , fz(x 2 −y 2 ) , fx(x 2 −3y 2 ) and f y(3x 2 −y 2 ) . (Reproduces with permission from S.A. Cotton, Lanthanides and Actinides, Macmillan, 1991). their isolation from the influence of the ligands, crystal-field effects are very small (and can be regarded as a perturbation on the free-ion states) and thus electronic spectra and magnetic properties are essentially unaffected by environment. The ability to form π bonds is also absent, and thus there are none of the M=O or M≡N bonds found for transition metals (or, indeed, certain early actinides). The organometallic chemistry is appreciably different from that of transition metals, too. 2.4 The Lanthanide Contraction As the series La–Lu is traversed, there is a decrease in both the atomic radii and in the radii of the Ln3+ ions, more markedly at the start of the series. The 4f electrons are ‘inside’ the 5s and 5p electrons and are core-like in their behaviour, being shielded from the ligands, thus taking no part in bonding, and having spectroscopic and magnetic properties largely independent of environment. The 5s and 5p orbitals penetrate the 4f subshell and are not shielded from 12 The Lanthanides – Principles and Energetics increasing nuclear charge, and hence because of the increasing effective nuclear charge they contract as the atomic number increases. Some part (but only a small fraction) of this effect has also been ascribed to relativistic effects. The lanthanide contraction is sometimes spoken of as if it were unique. It is not, at least in the way that the term is usually used. Not only does a similar phenomenon take place with the actinides (and here relativistic effects are much more responsible) but contractions are similarly noticed on crossing the first and second long periods (Li–Ne; Na–Ar) not to mention the d-block transition series. However, as will be seen, because of a combination of circumstances, the lanthanides adopt primarily the (+3) oxidation state in their compounds, and therefore demonstrate the steady and subtle changes in properties in a way that is not observed in other blocks of elements. The lanthanide contraction has a knock-on effect in the elements in the 5d transition series. It would naturally be expected that the 5d elements would show a similar increase in size over the 4d transition elements to that which the 4d elements demonstrate over the 3d metals. However, it transpires that the ‘lanthanide contraction’ cancels this out, almost exactly, and this has pronounced effects on the chemistry, e.g. Pd resembling Pt rather than Ni, Hf is extremely similar to Zr. 2.5 Electron Configurations of the Lanthanide Elements and of Common Ions The electronic arrangements of the lanthanide atoms have already been mentioned (Section 2.1 and Table 2.1) where it can be seen that in general the ECs of the atoms are [Xe] 4f n 5d0 6s2 , the exceptions being La and Ce, where the 4f orbitals have not contracted sufficiently to bring the energy of the 4f electrons below that of the 5d electrons; Gd, where the effect of the half-filled 4f subshell dominates; and Lu, the 4f subshell having already been filled at Yb. In forming the ions, electrons are removed first from the 6s and 5d orbitals (rather reminiscent of the case of the transition metals, where they are removed from 4s before they are taken from 3d), so that all the Ln3+ ions have [Xe] 4f n arrangements. 2.6 Patterns in Ionization Energies The values of the first four ionization energies for the lanthanides (and yttrium) are listed in Table 2.2. As usual, for a particular element, I4 > I3 > I2 > I1 , as the electron being removed is being taken from an ion with an increasingly positive charge, affording greater electrostatic attraction. Yttrium has greater ionization energies than the lanthanides as there is one fewer filled shell, and decreased distance effects outweigh the effect of the reduced nuclear charge. There is in general a tendency for ionization energies to increase on crossing the series but it is irregular. The low I3 values for gadolinium and lutetium, where the one electron removed comes from a d orbital, not an f orbital, and the high I3 value for Eu and Yb show some correlation with the stabilizing effects of half-filled and filled f sub shells. Results can be presented diagrammatically as cumulative ionization energies (Figure 2.3). These diagrams demonstrate at a glance the effect of the alternately high and low values of I3 for the elements Eu and Gd, influencing the respective stabilities of the +2 and +3 states of these elements, similarly illustrated by the neighbours Yb and Lu. The low I4 value for Ce (and to some extent for Pr and Tb) can be correlated with the accessibility of the Ce4+ ion. Patterns in Ionization Energies 13 Table 2.2 Ionization Energies (kJ/mole) La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu Y I1 I2 I3 I4 I1 + I2 I1 + I2 + I3 I1 + I2 + I3 + I4 538 527 523 529 536 543 546 593 564 572 581 589 597 603 523 616 1067 1047 1018 1035 1052 1068 1085 1167 1112 1126 1139 1151 1163 1176 1340 1181 1850 1949 2086 2130 2150 2260 2404 1990 2114 2200 2204 2194 2285 2415 2033 1980 4819 3547 3761 3899 3970 3990 4110 4250 3839 4001 4110 4115 4119 4220 4360 5963 1605 1574 1541 1564 1588 1611 1631 1760 1676 1698 1720 1740 1760 1779 1863 1797 3455 3523 3627 3694 3738 3871 4035 3750 3790 3898 3924 3934 4045 4194 3896 3777 8274 7070 7388 7593 7708 7990 8145 8000 7629 7899 8034 8049 8164 8414 8256 9740 Figure 2.3 Cumulative ionization energies across the lanthanide Series (reproduced by permission of Macmillan from S.A. Cotton, Lanthanides and Actinides, Macmillan, 1991). 14 The Lanthanides – Principles and Energetics 2.7 Atomic and Ionic Radii These are listed in Table 2.3 and shown in Figure 2.4. It will be seen that the atomic radii exhibit a smooth trend across the series with the exception of the elements europium and ytterbium. Otherwise the lanthanides have atomic radii intermediate between those of barium in Group 2A and hafnium in Group 4A, as expected if they are represented as Ln3+ (e− )3 . Because the screening ability of the f electrons is poor, the effective nuclear charge experienced by the outer electrons increases with increasing atomic number, so that the atomic radius would be expected to decrease, as is observed. Eu and Yb are exceptions to this; because of the tendency of these elements to adopt the (+2) state, they have the structure [Ln2+ (e− )2 ] with consequently greater radii, rather similar to barium. In contrast, the ionic radii of the Ln3+ ions exhibit a smooth decrease as the series is crossed. The patterns in radii exemplify a principle enunciated by D.A. Johnson: ‘The lanthanide elements behave similarly in reactions in which the 4f electrons are conserved, and very differently in reactions in which the number of 4f electrons change’ (J. Chem. Educ., 1980, 57, 475). Table 2.3 Atomic and ionic radii of the lanthanides (pm) Ba La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu Hf 217.3 187.7 182.5 182.8 182.1 181.0 180.2 204.2 180.2 178.2 177.3 176.6 175.7 174.6 194.0 173.4 156.4 La3+ Ce3+ Pr3+ Nd3+ Pm3+ Sm3+ Eu3+ Gd3+ Tb3+ Dy3+ Ho3+ Er3+ Tm3+ Yb3+ Lu3+ Y3+ 103.2 101.0 99.0 98.3 97.0 95.8 94.7 93.8 92.3 91.2 90.1 89.0 88.0 86.8 86.1 90.0 2.8 Patterns in Hydration Energies (Enthalpies) for the Lanthanide Ions Table 2.4 shows the hydration energies (enthalpies) for all the 3+ lanthanide ions, and also values for the stablest ions in other oxidation states. Hydration energies fall into a pattern Ln4+ > Ln3+ > Ln2+ , which can simply be explained on the basis of electrostatic attraction, 210 Radius (pm) 190 170 Ln3+ radius/pm metalic radius/pm 150 130 110 90 La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu Figure 2.4 Metallic and ionic radii across the lanthanide series. Enthalpy Changes for the Formation of Simple Lanthanide Compounds 15 since the ions with a larger charge have a greater charge density. The hydration energies for the Ln3+ ions show a pattern of smooth increase with increasing atomic number, as the ions becomes smaller and their attraction for water molecules increases. This is another example of the principle enunciated by Johnson. 2.9 Enthalpy Changes for the Formation of Simple Lanthanide Compounds Observed patterns of chemical behaviour can sometimes appear confusing or sometimes just be encapsulated in rules [e.g. Ce has a stable (+4) oxidation state]. They can frequently be explained in terms of the energy changes involved in the processes. 2.9.1 Stability of Tetrahalides Among the fluorides, lanthanum forms only LaF3 , whilst cerium forms CeF3 and CeF4 . (Tetrafluorides are also known for Pr and Tb, see Section 3.4). Why do neighbouring metals behave so differently? First, examine the energetics of formation of LaF3 using a Born– Haber cycle. H (kJ/mol) La(s) → La(g) +402 3+ La(g) → La (g) +538 + 1067 + 1850 3/2F2 (g) → 3F (g) +252 3F(g) + 3 e → 3F− (g) −984 La3+ (g) + 3F− (g) → LaF3 −4857 Thus La(s) + 3/2F2 (g) → LaF3 H = +402 + (538 + 1067 + 1850) + 252 − 984 − 4857 = −1732 kJ/mol Calculated and observed enthalpies of formation for LnXn (n = 2–4) are given in Table 2.5. The same method can be used to calculate Hf for LaF4 , making the assumption that the lattice energy is the same as that for CeF4 (−8391 kJ/mol), and that I4 for lanthanum is +4819 kJ/mol. In this case, Hf (LaF4 ) = −691 kJ/mol. Similarly, using the same method, Hf for LaF2 can be calculated as H = −831 kJ/mol (assuming that the lattice energy is the same as for BaF2 (−2350 kJ/mol). This poses the question: if Hf for LaF2 and for LaF4 are −831 and −691 kJ/mol respectively, why can’t these compounds be isolated? Apart from the oversimplification of using H rather than G values as an index of stability, what the preceding calculations have done is to indicate that these compounds are stable with respect to the elements, and no other decomposition pathways have been considered. Table 2.4 Enthalpies of hydration of the lanthanide ions (values given as −H hydr/kj mol−1 La3+ Ce3+ Pr3+ Nd3+ Pm3+ Sm3+ Eu3+ Gd3+ 3278 3326 3373 3403 Ce4+ 6309 3427 3449 Sm2+ 1444 3501 3517 Eu2+ 1458 Tb3+ Dy3+ Ho3+ Er3+ Tm3+ Yb3+ Lu3+ Y3+ 3559 3567 3623 3637 3664 3706 Yb2+ 1594 3722 3583 16 The Lanthanides – Principles and Energetics Table 2.5 Enthalpies of formation of lanthanide halidesa La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu LnF2 LnF3 LnF4 LnCI2 LnCI3 LnCI4 LnBr2 LnBr3 LnI2 LnI3 880 950 1050 1050 1080 1160 1188 870 980 1050 1040 1020 1090 1172 790 1732 1733 1712 1661 1700 1669 1584 1699 1707 1678 1698 1699 1689 1570 1640 600 1946 1690 1500 1500 1450 1290 1310 1742 1540 1500 1510 1500 1250 1210 520 580 700 707 720 820 824 500 600 693 660 640 709 799 420 1073 1058 1059 1042 1040 1040 1062 937 1008 1007 990 995 995 960 986 −480 820 630 490 430 400 230 180 600 420 350 360 360 250 160 430 490 610 630 620 720 760 400 500 590 580 560 640 710 330 907 890 891 873 850 857 799 829 850 831 830 836 840 800 850 320 380 490 510 510 600 630 290 390 470 450 420 500 580 190 699 686 678 665 640 640 540 619 598 603 594 586 582 561 556 are quoted as −Hf (kJ/mol). Experimental values in bold. Values taken from D.W. Smith, J. Chem. Educ., 1986, 63, 228. a Values LaF4 might decompose thus : LaF4 → LaF3 + 1/2 F2 Applying Hess’s Law to this in the form Hreaction = gH f (products) − H f (reactants) Hreaction = −1732 − (−691 + 0) = −1041 kJ/mol This decomposition is thus thermodynamically favourable (especially as it would be favoured on entropy grounds too, with the formation of fluorine gas) LaF2 might decompose by disproportionation: 3LaF2 → 2LaF3 + La Using Hf for LaF2 and for LaF3 (−831 and −1732 kJ/mol respectively), H for this decomposition reaction can be calculated as −971 kJ/mol. This is again a very exothermic process, indicating that LaF2 is likely to be unstable (this is analogous to the reason for the non-existence of MgCl, as disproportionation into Mg and MgCl2 is favoured, as the reader may be aware). The reason for this is that, although I3 for lanthanum is large (and endothermic), it is more than compensated for by the higher lattice energy for LaF3 compared with the value for LaF2 . Since both CeF3 and CeF4 are isolable, what makes the difference here? If similar calculations are carried out (assuming lattice energies for CeF3 and CeF4 of −4915 and −8391 kJ/mol respectively, and an enthalpy of atomization of 398 kJ/mol for Ce (see Table 2.6), and using the same enthalpy of atomization and electron affinity for fluorine as in the lanthanum examples), Hf for CeF3 can be calculated as −1726 kJ/mol and Hf for CeF4 as −1899 kJ/mol. The discrepancy between Hf for CeF3 and CeF4 is much smaller than is the case for lanthanum. In fact, for the decomposition reaction CeF4 → geF3 + 1/2 F2 Enthalpy Changes for the Formation of Simple Lanthanide Compounds 17 Table 2.6 Enthalpies of atomization of the lanthanides (kJ/mol) Ba 150.9 La Ce Pr Nd 402.1 398 357 328 Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu Hf 164.8 176 301 391 293 303 280 247 159 428.0 570.7 H = −1726 − (−1899) = +173 kJ/mol, making the decomposition of CeF4 relatively unfavourable. CeF4 is stable, certainly at ambient temperatures. If the values of the parameters used in the calculation are compared, the determining factor is the much lower value of I4 for cerium (3547 kJ/mol compared with 4819 kJ/mol for La) due to the fact that the fourth electron is being removed from a different shell (nearer the nucleus) in the case of lanthanum. Other tetrahalides do not exist. Thus, though both CeCl3 and salts of the [CeCl6 ]2− ion can be isolated, CeCl4 cannot be made. The reasons for this are those that enable fluorine to support high oxidation states (see Question 2.4). Similar factors indicate that tetrabromides and tetraiodides are much less likely to be isolated. 2.9.2 Stability of Dihalides The stability of the dihalides can be explained in a similar way. As already noted, LaF2 does not exist. When might dihalides be expected? One way in which dihalides can decompose is disproportionation 3 LnX2 → 2 LnX3 + Ln though a number of dihalides (particularly dichlorides) have been made by the reverse reproportionation, by heating the mixture to a high temperature and rapidly quenching (Section 3.5.1.) 2 LnX3 + Ln → 3 LnX2 It can be shown, using Hess’s Law, that the disproportionation will be exothermic unless: Hf LnX3 /Hf LnX2 < 1.5 The disproportionation can be broken down into individual components (Figure 2.5) H1 = H2 + H3 + H4 H1 = [−3Hlatt (LnX2 )] + {2 I3 (Ln) − [I1 (Ln) + I2 (Ln)]} + [−Hat (Ln) + 2 Hlatt (LnX3 )] H1 = {2 I3 (Ln) − [I1 (Ln) + I2 (Ln)]} + 2 Hlatt (LnX3 ) − 3 Hlatt (LnX2 ) − Hat (Ln) 3 LnX2(s) ∆H1 2 LnX3(s) + Ln(s) ∆H2 2+ ∆H4 − 3 Ln (g) + 6X (g) Figure 2.5 Disproportion of lanthanide dihalides. ∆H3 3+ − 2 Ln (g) + 6X (g) + Ln(g) 18 The Lanthanides – Principles and Energetics These equations can be used qualitatively, first to suggest for which lanthanides the halides LnX2 are most likely to be stable. For the disproportionation process to be more likely to be endothermic (i.e. stabilizing the +2 state), the preceding equation suggests that high values of I3 are favourable. Study of Table 2.2 shows that this is more likely to be associated with Eu, Sm and Yb. Iodide is the halide most often found in low oxidation state halides. In the case of LnI2 , the large size of the iodide ion will reduce the lattice energy for both LnI2 and LnI3 so that the difference in lattice enthalpy will become less significant, favouring LnI2 . The dihalide will also be favoured by a low Hat (Ln), again associated with the lanthanides most often found in the +2 state, Eu and Yb (see Table 2.6). Applying the Hf LnX3 /Hf LnX2 <1.5 criterion, and using data in Table 2.5, the halides LnX2 are most likely to be stable are predicted to be LnF2 (Ln = Sm, Eu, Yb); LnCl2 (Ln = Nd, Pm, Sm, Eu, Dy, Tm, Yb); LnY2 (Y = Br, I; Ln = Pr, Nd, Pm, Sm, Eu, Dy, Ho, Er, Tm, Yb). The known dihalides are listed in Table 3.2. There is a reasonably good correlation, given that the Pm dihalides have not been investigated on account of promethium’s short half-life. Some dihalides listed are ‘metallic’ and are not covered by this argument. 2.9.3 Stability of Aqua Ions Since Ln3+ (aq) is the most stable aqua ion, then both of the following processes are favoured. Ln2+ (aq) + H+ (aq) → Ln3+ (aq) + 1/2 H2 (g) and 2 Ln4+ (aq) + H2 O(aq) → 2 Ln3+ (aq) + 2 H+ (aq) + 1/2 O2 (g) In other words, Ln2+ ions tend to reduce water and Ln4+ ions tend to oxidize it. We can examine the stability of the Ln2+ ion using a treatment similar to the one just employed (Figure 2.6), with Hox (Ln2+ ) = I3 + [Hhydr [Ln3+ (aq)] − Hhydr [Ln2+ (aq)] − 439 kJ/mol. When is it most likely that Ln2+ (aq) ions will be stable? For the first of the two reactions above to be favoured, the single factor that will help make H positive is a high value of I3 . Less important would be the size of the ions, as this could affect the hydration enthalpies; the difference between the hydration enthalpies will be less, the larger the lanthanide ions. Substituting into the above equation, we can investigate the relative stabilities of La2+ (aq) and Eu2+ (aq), making use of ionization energies from ∆ Hhydr (Ln2+ ) Ln2+(g) + Ln2+(aq) + H (aq) I3 Ln3+(g) Figure 2.6 Oxidation of Ln2+ (aq). ∆HOX(Ln2+) + ∆HH ∆ Hhydr (Ln3+) Ln3+(aq) + 1/2H2(g) Patterns in Redox Potentials 19 Table 2.2 and enthalpies of hydration found in Table 2.4, also assuming Hhydr [(La 2+ (aq)] = −1327 kJ/mol: For La2+ : Hox (La2+ ) = I3 + {Hhydr [La3+ (aq)] − Hhydr [La2+ (aq)]} − 439 = 1850 + [−3278 − (−1327)] − 439 = −540 kJ/mol For Eu2+ : Hox (Eu2+ ) = I3 + {Hhydr [Eu3+ (aq)] − Hhydr [Eu2+ (aq)]} − 439 = 2404 + [−3501 − (−1458)] − 439 = −78 kJ/mol The large exothermic value for Hox (La2+ ) indicates that it is not likely to exist in aqueous solution. The Eu2+ (aq) ion is known to have a short lifetime in water, even though H is negative for the oxidation process, so the activation energy for oxidation may be rather high. 2.10 Patterns in Redox Potentials Known and estimated values are listed in Table 2.7. The values for the reduction potential for Ln3+ + 3 e− → Ln are very consistent, with slight irregularities at Eu and Yb. The potential largely depends upon three processes: Ln(s) → Ln(g) Hat (Ln) Ln(g) → Ln (g) + 3 e− 3+ Ln (g) → Ln (aq) 3+ 3+ I 1 + I 2 + I3 Hhydr (Ln3+ ) The first two of these are endothermic and the third exothermic; overall H is the difference between two large quantities. It remains fairly constant across the series, apart from Eu and Yb, with values of 608 (La); 712 (Eu); 630 (Gd); 613 (Tm); 644 (Yb) and 593 (Lu) kJ/mol being representative, the values for Eu and Yb resulting from the high I3 values. The very negative value for the reduction potential is expected for such reactive metals (and also reflects the difficulty in isolating them). The potentials for the Ln3+ + e− → Ln2+ process reflect the stability of the +2 state. Since I3 relates to H for the process, and G = −nFE, a relationship between these is unsurprising. Similarly, the only potential for the Ln4+ + e− → Ln3+ process within reasonable range is that for Ce, and indicates that Ce4+ is the only ion in this state likely to be encountered in aqueous solution. Question 2.1 What is the trend in atomic radii, and that in ionic radii, for the lanthanides? What are the exceptions to this, and why? Answer 2.1 The structure of metals is usually described as one in which metal ions are surrounded by a ‘sea’ of delocalised outer-shell electrons. The greater the number of loosely held electrons, the stronger the metallic bonding and the smaller the atomic radius. If the Ce Pr b= in parentheses are estimated. in THF. a Values Ln3+ + 3e → Ln −2.37 −2.34 −2.35 Ln3+ + e → Ln2+ (−3.1) (−3.2) (−2.7) 1.70 (3.4) Ln4+ + e → Ln3+ La Pm −2.32 −2.29 −2.6b (−2.6) (4.6) (4.9) Nd Table 2.7 Redox potentials of the lanthanide ions (V)a −2.30 −1.55 (5.2) Sm Gd Tb −1.99 −2.29 −2.30 −0.34 (−3.9) (−3.7) (6.4) (7.9) (3.3) Eu Ho Er −2.29 −2.33 −2.31 −2.5b (−2.9) (−3.1) (5.0) (6.2) (6.1) Dy −2.31 −2.3b (6.1) Tm −2.22 −1.05 (7.1) Yb Y (8.5) −2.30 −2.37 Lu Patterns in Redox Potentials 21 lanthanides are represented as Ln3+ (e− )3 , then the atomic radii would be expected to fall between those of barium [Ba2+ (e− )2 ] and hafnium [Hf 4+ (e− )4 ], as is generally observed. Question 2.2 Using Hess’s Law and the Hf for LaF2 (−831 kJ/mol) and for LaF3 (−1732 kJ/mol), calculate H for this decomposition reaction. 3 LaF2 → 2 LaF3 + La Answer 2.2 Hreaction = Hf (products) − Hf (reactants) Hreaction = 2 (−1732) − 3 (−831 + 0) = −971 kJ/mol Question 2.3 Assuming lattice energies for CeF3 and CeF4 of −4915 and −8391 kJ/mol respectively, and using the same enthalpy of atomization and electron affinity for fluorine as in the lanthanum examples (Section 2.9.1.), calculate Hf for CeF3 and CeF4 . Take an enthalpy of atomization of 398 kJ/mol for cerium (Table 2.6). Use ionization energies from Table 2.2. Answer 2.3 H (kJ/mol) +398 +527 + 1047 + 1949 +252 −984 −4915 Ce(s) → Ce (g) Ce(g) → Ce3+ (g) 3/2 F2 (g) → 3F(g) 3F(g) + 3e → 3F− (g) Ce3+ (g) + 3F− (g) → CeF3 (s) Thus Ce(s) + 3/2 F2 (g) → CeF3 (s) H = +398 + (527 + 1047 + 1949) + 252 − 984 − 4915 = −1726 kJ/mol H (kJ/mol) +398 +527 + 1047 + 1949 + 3547 +336 −1312 −8391 Ce(s) → Ce(g) Ce(g) → Ce4+ (g) 2F2 (g) → 4F (g) 4F(g) + 4e → 4F− (g) Ce4+ (g) + 4F− (g) → CeF4 (s) Thus Ce(s) + 2F2 (g) → CeF4 (s) H = +398 + (527 + 1047 + 1949 + 3547) + 336 − 1312 − 8391 = −1899 kJ/mol Question 2.4 Unlike CeF4 , CeCl4 does not exist, though CeCl3 does. Suggest why this might be. Values of Hf for CeCl3 and CeCl4 are −1058 and −820 (calculated) kJ/mol, respectively. Answer 2.4 Fluorine is well known to promote high oxidation states. Factors associated with this are the high lattice energies associated with the small fluoride ion, along with the very small F–F bond energy (due to non-bonding electron pair repulsions) as well as high bond energies involving fluorine (not relevant in this case). Because of the larger size of the Cl− ion, there is going to be much less difference between the lattice energies of CeCl3 and 22 The Lanthanides – Principles and Energetics =E = I3 0 1800 3 2000 2 2200 1 2400 E 0(V) 4 I3(kJ mol−1) Key: La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu Figure 2.7 Strong correlation between I3 and E ◦ for Ln3+ + e → Ln2+ (reproduced by permission of Macmillan from S.A. Cotton, Lanthanides and Actinides, 1991, p. 26.) CeCl4 , and this is probably the determining factor. The higher Hat (Cl) of 121.5 kJ/mol also mitigates against the formation of CeCl4 . Question 2.5 Applying the Hf LnX3 /Hf LnX2 < 1.5 criterion for stability of dihalides, and using data in Table 2.5, predict which the halides LnX2 are most likely to be exist. Answer 2.5 As in section 2.9.2 LnF2 (Ln = Sm, Eu, Yb); LnCl2 (Ln = Nd, Pm, Sm, Eu, Dy, Tm, Yb); LnY2 (Y = Br, I; Ln = Pr, Nd, Pm, Sm, Eu, Dy, Ho, Er, Tm, Yb) are predicted to be stable, in reasonably good agreement with the currently known facts. Exercise 2.6 Using the same horizontal axis (atomic numbers 57–71), plot values of (a) the third ionization enthalpy I3 and (b) the Ln3+ + e → Ln2+ reduction potential on the y axis (choose appropriate scales). Comment. Answer 2.6 There is a strong correlation between them, not surprisingly! See Figure 2.7. 3 The Lanthanide Elements and Simple Binary Compounds By the end of this chapter you should be able to: r know how to prepare lanthanide metals and simple binary compounds such as the halides, oxides, and hydrides; the principle of the lanthanide contraction to explain patterns in co-ordination number in these compounds; apply knowledge gained in the study of Chapter 2 to the compounds in unusual oxidation states; understand the uses of the metals and certain compounds in applications such as hydrogen storage and in superconductors. r apply r r 3.1 Introduction This chapter discusses the synthesis of the lanthanide metals, their properties, reactions, and uses. It also examines some of the most important binary compounds of the lanthanides, particularly the halides, which well illustrate patterns and trends in the lanthanide series. 3.2 The Elements 3.2.1 Properties The lanthanides are rather soft reactive silvery solids with a metallic appearance, which tend to tarnish on exposure to air. They react slowly with cold water and rapidly in dilute acid. They ignite in oxygen at around 150–200 ◦ C; similar reactions occur with the halogens, whilst they react on heating with many nonmetals such as hydrogen, sulfur, carbon, and nitrogen (above 1000 ◦ C). The metals are relatively high-melting and -boiling. Their physical properties usually show smooth transition across the series, except that discontinuities are often observed for the metals that have a stable +2 state, europium and ytterbium. Thus the atomic radii of europium and ytterbium are about 0.2 Å greater than might be predicted by interpolation from values for the flanking lanthanides (Figure 3.1). Similarly, Sm, Eu, and Yb have boiling points that are lower than those of the neighbouring metals (Figure 3.2). Lanthanide and Actinide Chemistry S. Cotton C 2006 John Wiley & Sons, Ltd. 24 The Lanthanide Elements and Simple Binary Compounds Figure 3.1 The atomic radii of the lanthanide metals (reproduced with permission from S.A. Cotton, Lanthanides and Actinides, Macmillan, 1991). 4000 Bp (K) 3000 2000 1000 La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu Figure 3.2 The boiling points of the lanthanide metals (reproduced with permission from S.A. Cotton, Lanthanides and Actinides, Macmillan, 1991). Assuming that one can represent the structure of a metal as a lattice of metal ions permeated by a sea of electrons, then metals like lanthanum can be shown as (Ln3+ ) (e− )3 ; however, metals based upon divalent ions (like Eu and Yb) would be (Ln2+ ) (e− )2 . The ions with the (+3) charge have a smaller radius, as the higher charge draws in the electrons more closely, and the stronger attraction means that it takes more energy to boil them (similarly, they would be predicted to have higher conductivities). 3.2.2 Synthesis The metals have similar reactivities to magnesium. This means that they cannot be extracted by methods like carbon reduction of the oxides; in practice, metallothermic reduction of the lanthanide fluoride or chloride with calcium at around 1450 ◦ C is used. The product Binary Compounds 25 is an alloy of (excess of) calcium and the lanthanide, from which the calcium can be distilled. 2 LnF3 + 3 Ca → 2 Ln + 3 CaF2 The reduction is carried out under an atmosphere of argon, not nitrogen. The above method is not suitable for obtaining the metals with a stable +2 state, which are only reduced as far as the difluoride (Eu, Yb, Sm). The lanthanide can be removed by distillation. 2 La + M2 O3 2 M + La2 O3 The ‘divalent’ metals Sm, Eu, and Yb have boiling points of 1791, 1597 and 1193 ◦ C respectively, much lower than that of La (3457 ◦ C), so that on heating they are distilled off, their volatility meaning that their removal from the mixture will displace the equilibrium to the right, so the reaction will proceed to completion. 3.2.3 Alloys and Uses of the Metals Mischmetal is the lanthanide alloy with the longest history. It is a mixture of the lighter lanthanides, cerium in particular, which is manufactured from an ‘unseparated’ mixture of the oxides. This is first converted into a mixture of the anhydrous chlorides, which is then electrolysed using a graphite anode and iron cathode at around 820 ◦ C to afford the mixture of metals. This is used mainly as an alloy with iron for the desulfurization and deoxidation of steels and more familiarly in cigarette lighter flints. SmCo5 and other alloys of these metals have been used to make extremely strong permanent magnets. LaNi5 has been widely examined as a material for hydrogen storage, with applications in fuel cells, catalytic hydrogenation, and removal of hydrogen from gas mixtures, as it rapidly absorbs hydrogen at room temperature to afford compositions up to LaNi5 H6 ; the hydrogen is given up quickly at 140 ◦ C. The most important application lies in rechargeable batteries for PCs (see Section 3.10). 3.3 Binary Compounds 3.3.1 Trihalides Most halides are LnX3 , but a number of LnX2 are known, as are a handful of tetrafluorides. Syntheses of the Trihalides Although the halides can be obtained as hydrates from reaction of the metal oxides or carbonates with aqueous acids, these hydrates are hydrolysed on heating to the oxyhalide and thus the anhydrous halides (which are themselves deliquescent) cannot be made that way. LnX3 + H2 O → LnOX + HX The fluorides are obtained as insoluble hydrates LnF3 .0.5H2 O by precipitation and the hydrates dehydrated by heating in a current of anhydrous HF gas (or in vacuo). LnF3 .0.5H2 O → LnF3 + 0.5 H2 O 26 The Lanthanide Elements and Simple Binary Compounds Otherwise the anhydrous halides can generally be made by heating the metal with the halogen (except for EuI3 ) or gaseous HCl. 2 Ln + 3 X2 = 2 LnX3 Another method for the chlorides involves refluxing the hydrated chlorides with thionyl dichloride (SOCl2 ) for a few hours; an advantage of this method is that the other reaction products are gaseous SO2 and HCl. LnCl3 .xH2 O + x SOCl2 → LnCl3 + x SO2 + 2x HCl A route that works well in practice involves thermal decomposition of ammonium halogenometallates. Adding ammonium chloride to a solution of the metal oxides in hydrochloric acid, followed by evaporation gives halogenometallate salts. These can be dehydrated by heating with excess of ammonium chloride in a stream of gaseous HCl, the resulting anhydrous salt being decomposed by heating in vacuo at about 300 ◦ C. Ln2 O3 + 9 NH4 Cl + 3 HCl → 2 (NH4 )3 LnCl6 + 3 NH3 + 3 H2 O 2 (NH4 )3 LnCl6 → 2 LnCl3 + 6 NH4 Cl The anhydrous halides can be purified by sublimation in vacuo, but owing to a tendency to react with silica, contact with hot glass, thereby forming the oxyhalide, should be avoided. Structures of the Trihalides The lanthanide trihalides demonstrate very clearly the effect of varying the cation and anion radii upon the structure type adopted (Table 3.1). The early lanthanide fluorides adopt the ‘LaF3 ’ structure (Figure 3.3) based on a metal ion surrounded by a trigonal prism of fluorides with two additional capping fluorides, giving 11 coordination (9 + 2), whilst the later fluorides have the ‘YF3 ’ structure. This is based on tricapped trigonal prismatic 9 coordination, in which the prism is somewhat distorted. The structure adopted by UCl3 and several of the lanthanide halides is again a tricapped Table 3.1 Structures of the lanthanide trihalides La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu Y F Cl Br I LaF3 (11) LaF3 (11) LaF3 (11) LaF3 (11) LaF3 (11) YF3 (9) YF3 (9) YF3 (9) YF3 (9) YF3 (9) YF3 (9) YF3 (9) YF3 (9) YF3 (9) YF3 (9) YF3 (9) UCl3 (9) UCl3 (9) UCl3 (9) UCl3 (9) UCl3 (9) UCl3 (9) UCl3 (9) UCl3 (9) PuBr3 (8) AlCl3 (6) AlCl3 (6) AlCl3 (6) AlCl3 (6) AlCl3 (6) AlCl3 (6) AlCl3 (6) UCl3 (9) UCl3 (9) UCl3 (9) PuBr3 (8) PuBr3 (8) PuBr3 (8) PuBr3 (8) FeCl3 (6) FeCl3 (6) FeCl3 (6) FeCl3 (6) FeCl3 (6) FeCl3 (6) FeCl3 (6) FeCl3 (6) FeCl3 (6) PuBr3 (8) PuBr3 (8) PuBr3 (8) PuBr3 (8) PuBr3 (8) FeCl3 (6) Values in parentheses indicate the coordination number. FeCl3 FeCl3 FeCl3 FeCl3 FeCl3 FeCl3 FeCl3 FeCl3 FeCl3 (6) (6) (6) (6) (6) (6) (6) (6) (6) Binary Compounds 27 Figure 3.3 The LaF3 structure (reproduced by permission of Macmillan from S.A. Cotton, Lanthanides and Actinides, 1991, p. 42). Figure 3.4 The structure of UCl3 and certain LnCl3 [reproduced in modified form, from R.B. King (ed.), Encyclopedia of Inorganic Chemistry, 1st edition, Wiley, Chichester; 1994]. trigonal prism (Figure 3.4); if one of the face-capping halogens is removed from the ‘UCl3 ’ structure, the 8 coordinate PuBr3 structure is generated. Finally, the AlCl3 and BiI3 (also sometimes referred to as ‘FeCl3 ’) structures both have octahedral six-coordination. The trends are similar to those found in halides of the Group I and Group II metals and are explained on an ionic packing model; more anions can be packed round the large lanthanide ions early in the series than around the smaller, later ones; similarly, more of the small fluoride ions can be packed round a given lanthanide ion than is the case with the much larger iodide anion. Properties of the Trihalides The trihalides are high-melting solids, with many uses in synthetic chemistry, though their insolubility in some organic solvents means that complexes, such as those with thf (Section 4.3.3), are often preferred. 3.3.2 Tetrahalides Only the fluorides of Ce, Pr, and Tb exist, the three lanthanides with the most stable (+4) oxidation state. Fluorine is most likely to support a high oxidation state, and even though salts of ions like [CeCl6 ]2− are known, the binary chloride has not been made. CeF4 can be crystallized from aqueous solution as a monohydrate. Anhydrous LnF4 (Ln = Ce, Pr, Tb) 28 The Lanthanide Elements and Simple Binary Compounds can be made by fluorination of the trifluoride or, in the case of Ce, by fluorination of metallic Ce or CeCl3 . All three tetrafluorides have the MF4 structure with dodecahedral eight coordination. Factors that favour formation of a tetrafluoride include a low value of I4 for the metal and a high lattice energy (see Section 2.9.1). This is most likely to be found with the smallest halide ion, fluoride. The low bond energy of F2 is also a supporting factor. 3.3.3 Dihalides These are most common for metals with a stable (+2) state, such as Eu, Yb, and Sm. As would be expected, they most often occur for the iodide ion, the best reducing agent. A number of dihalides are known for other metals, though some of these do not actually involve the (+2) oxidation state. Synthetic Routes These compounds are usually made by reduction using hydrogen (e.g. EuX2 , YbX2 , or SmI2 ) or reproportionation. 2 EuCl3 + H2 → 2 EuCl2 + 2 HCl 2 DyCl3 + Dy → 3 DyCl2 In a few cases, thermal decomposition is applicable [LnI2 (Ln = Sm, Yb); EuBr2 ]. 2 YbI3 → 2 YbI2 + I2 Another method, used especially for diiodides, involves heating the metal with HgX2 . Tm + HgI2 → TmI2 + Hg The known dihalides are listed in Table 3.2, together with an indication of the structure and coordination number. The iodides of Nd, Sm, Eu, Dy, Tm, and Yb are definitely compounds of the +2 ions, with salt-like properties, are insulators, and have magnetic and spectroscopic properties expected Table 3.2 Structures of the lanthanide dihalides F La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu a Cl Br I a MoSi2 (8) MoSi2 (8) a MoSi2 (8) SrBr2 (7,8) a CaF2 (8) CaF2 (8) CaF2 (8) PbCl2 (9) PbCl2 (9) PbCl2 (9) PbCl2 (9) PbCl2 (9); SrBr2 (7,8) SrBr2 (7,8) EuI2 (7) EuI2 (7) a MoSi2 (8) SrBr2 (7,8) SrI2 (7) CdCl2 (6) SrI2 (7) SrI2 (7) SrI2 (7) SrI2 (7); CaCl2 (6) CdI2 (6) CdI2 (6) LnIII compounds with the structure M3+ (I− )2 (e− ). Binary Compounds 29 for the M2+ ions (thus electronic spectra of EuX2 resemble those of the isoelectronic Gd3+ ions). SmI2 is proving an important reagent in synthetic organic chemistry (Section 8.3). Several of the diiodides, however, have a metallic sheen and are very good conductors of electricity, such as LaI2 , CeI2 , PrI2 , and GdI2 . Since they are good conductors in the solid state, the presence of delocalized electrons is indicated. M3+ (I− )2 (e− ) is a likely structure. None of these metals exhibits a stable +2 state in any of its compounds and consideration of factors such as their ionization energies suggest they are unlikely to form stable compounds in this state (see Section 2.9.2). 3.3.4 Oxides Oxides M2 O3 As already mentioned, the metals burn easily, rather like Group 2 metals, forming oxides (but with some nitride). The oxides are therefore best made by thermal decomposition of compounds like the nitrate or carbonate. 4 Ln(NO3 )3 → 2 Ln2 O3 + 12 NO2 + 3 O2 Most lanthanides form Ln2 O3 , but those metals with accessible +4 and +2 oxidation states can afford other stoichiometries. These may be turned into Ln2 O3 by synthesis under a reducing atmosphere. CeO2 can be reduced to Ce2 O3 using hydrogen. The oxides are somewhat basic and absorb CO2 from the atmosphere, forming carbonates, and water vapour, forming hydroxides. As expected they dissolve in acid, forming salts, and are convenient starting materials for the synthesis of lanthanide salts, including the hydrated halides. The sequioxides Ln2 O3 adopt three structures, depending upon the temperature and upon the lanthanide involved. At room temperature, La2 O3 to Sm2 O3 inclusive adopt the A-type structure, which has capped octahedral 7 coordination of the lanthanide. The B-type structure tends to be exhibited by La2 O3 to Sm2 O3 at higher temperatures and has three different lanthanide sites, one with distorted 6 coordination and the others with face-capped octahedral 7 coordination. The C-type structure is followed by heavier metals and has 6 coordination, severely distorted from an octahedron. Oxides MO2 Cerium, having the stablest (+4) state, is the only metal to form a stoichiometric oxide in this state, CeO2 , which has the fluorite structure. It is white when pure, but even slightly impure specimens tend to be yellow. It can be made by burning cerium or heating salts like cerium(III) nitrate strongly in air. It is basic and dissolves with some difficulty in acid, forming Ce4+ (aq), which can be isolated as salts such as the nitrate Ce(NO3 )4 .5H2 O. Uses of CeO2 include self-cleaning ovens and as an oxidation catalyst in catalytic converters. Praseodymium and terbium form higher oxides, of which a number of phases are known between Ln2 O3 and LnO2 . Ignition of praseodymium nitrate leads to Pr2 O3 but further heating in an oxygen atmosphere gives Pr6 O11 or even PrO2 ; terbium similarly yields Tb4 O7 and TbO2 . Oxides MO Reduction of Eu2 O3 with Eu above 800 ◦ C (comproportionation) gives EuO. Eu2 O3 + Eu → 3 EuO 30 The Lanthanide Elements and Simple Binary Compounds This compound and the similar YbO have salt-like (NaCl) structures and are genuine LnII compounds, being insulators. Similar comproportionation methods using high pressures have been used to obtain SmO and NdO; these are shiny conducting solids, probably containing Ln3+ ions. Oxide Superconductors Superconductivity is the phenomenon in which a material conducts electricity with virtually zero resistance. For many years until the mid 1980s, the highest temperatures available were around 20 K and required expensive liquid helium (or hydrogen) coolant. If hightemperature superconductors can be made, this has obvious application in area such as power transmission. In 1986, Bednorz and Muller reported that La1.8 Sr0.2 CuO4 was a superconductor up to 38 K. This sparked intense world-wide activity, and the following year Wu, Chu and others reported that YBa2 Cu3 O7−δ (0 ≤ δ ≤ 1) had a Tc (superconducting transition temperature) of 92 K, bringing the phenomenon into the liquid nitrogen range. Steady though less spectacular advances have produced materials like HgBa2 Ca2 Cu3 O8 (Tc = 133K). The structure of YBa2 Cu3 O7 is based on an oxygen-deficient layered perovskite structure and features two types of copper environment (Figure. 3.5), formally containing Cu2+ and Cu3+ , with both square planar and square pyramidal coordination. Removing the shaded oxygens results in phases down to the semiconductor YBa2 Cu3 O6 . The mechanism of superconductivity is still debated, but one theory suggests that it involves an electron passing through the lattice distorting it in such a way that a second electron follows closely in its wake Figure 3.5 The structure of YBa2 Cu3 O7 ; the shaded atoms are those removed to create oxygen-deficient phases up to YBa2 Cu3 O6 (reproduced by permission of Macmillan from S.A. Cotton, Lanthanides and Actinides, 1991, p. 47). Hydrides 31 with no hindrance to its passing, the electron pair being known as a ‘Cooper pair’. Despite intense activity over the last 15 years no truly commercial material has yet emerged, due to the intrinsically brittle nature of the oxide ceramic materials making fabrication into wires impossible. Current thinking favours making thin films, though with present technology they will only have low current-carrying capacity. 3.4 Borides A number of stoichiometries obtain, such as LnB2 , LnB4 , LnB6 , LnB12 , and LnB66 . The most important are LnB6 . Borides are obtained by heating the elements together at 2000 ◦ C or by heating the lanthanide oxide with boron or born carbide at 1800 ◦ C. They are extremely unreactive towards acids, alkalis, and other chemicals, and are metallic conductors. They contain a continuous three-dimensional framework of (B6 )2− octahedra interspersed with Ln3+ ions, indicating an electronic structure (Ln3+ ) (B6 )2− (e− ) in most cases, though EuB6 and YbB6 are insulators, suggesting them to be (Ln2+ ) (B6 )2− . Because of its high thermal stability and melting point (2400 ◦ C), and its metal-like electrical and thermal conductivity, LaB6 is an important thermionic emitter material for the cathodes of electron guns. Mixed boride materials are also of commercial importance. Nd–Fe–B alloys, with compositions such as Nd2 Fe14 B, are the strongest permanent magnet materials available today. Other important compounds include lanthanide rhodium boride low-temperature superconductors (such as ErRh4 B4 ), part of a wider LnM4 B4 family (M = transition metal). 3.5 Carbides The lanthanides form carbides with a range of compositions, notably LnC2 , but additionally Ln2 C3 , LnC, Ln2 C, and Ln3 C phases are known (depending upon the lanthanide in question and the conditions of synthesis). LnC2 adopt the CaC2 structure, containing isolated C2− 2 ions. 3.6 Nitrides These can be made by direct synthesis from the elements at 1000 ◦ C. They have the NaCl structure and can be hydrolysed to NH3 . 3.7 Hydrides Ternary hydrides such as LaNi5 H6 have attracted attention as materials for electrodes in fuel cells and for gas storage. Nickel/metal hydride batteries for notebook PCs are a less toxic alternative than Ni/Cd batteries; when charging, hydrogen generated at the negative electrode enters the lattice of the La/Ni alloy (in practice a material such as La0.8 Nd0.2 Ni2.5 Co2.4 Si0.1 is used to improve corrosion resistance, storage capacity, discharge rate, etc.). The overall cell reaction is: LaNi5 + 6 Ni(OH)2 LaNi5 H6 + 6 NiOOH 32 The Lanthanide Elements and Simple Binary Compounds The lanthanides also form simple binary hydrides on combination of the elements at about 300 ◦ C. These compounds have ideal compositions of MH2 and MH3 , but are frequently non-stoichiometric. Thus lutetium forms phases with ranges LuH1.83 to LuH2.23 and LuH2.78 to LuH3.00 . Reaction of ytterbium with hydrogen under pressure gives YbH2.67 . MH3 are obtained only at higher gas pressures, whilst europium, the lanthanide with the most stable +2 state, forms only EuH2 . The dihydrides are generally good electrical conductors, and thus are thought to be M3+ (H− )2 (e− ), whilst the trihydrides are salt-like nonconductors believed to be M3+ (H− )3 . The hydrides are reactive solids, owing to the presence of the easily hydrolysed H− ion. 3.8 Sulfides These are quite important compounds. A number of stoichiometries exist, the most important being Ln2 S3 . These can be made by direct synthesis, heating the elements together, or by passing H2 S over heated LnCl3 . Eu2 S3 cannot be prepared by this latter route. A range of compositions between Ln2 S3 and Ln3 S4 , the latter formed when Ln2 S3 lose sulfur on heating, generally occurs. Ln2 S3 are insulators, genuine Ln(III) compounds, but Ln3 S4 (having the Th3 P4 structure) are more complex. Some, like Ce3 S4 , are metallic conductors and are thus (Ln3+ )3 (S2− )4 (e− ); others, like Eu3 S4 and Sm3 S4 , are semiconductors and may be (Ln2+ )(Ln3+ )2 (S2− )4 . The structures of Ln2 S3 fall into a pattern. La2 S3 to Dy2 S3 adopt the ‘Gd2 S3 ’ structure with 7-coordinate lanthanide ions; Dy2 S3 to Tm2 S3 have the ‘Ho2 S3 ’ structure with 6- and 7-coordinate lanthanides, whilst Yb2 S3 and Lu2 S3 have the corundum structure with just 6 coordination. Monsulfides MS are formed by direct combination. They adopt the NaCl structure but have a variety of bonding types. YbS and EuS are genuine Ln2+ S2− but CeS exhibits the magnetic properties expected for Ce3+ and has a bronze metallic lustre as well, so it thought to be (Ce3+ )(S2− )(e− ). This substance not only has electrical conductivity in the metallic region but also can be machined like a metal too. SmS has some unusual properties; it is usually obtained as a black semiconducting phase, Sm2+ S2− , but can reportedly be turned into a golden metallic phase by the action of pressure, polishing or even scratching on single crystals. Oxysulfides are also rather important. Y2 O2 S is used as a host material