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

Only $11.99/month after trial. Cancel anytime.

Alpha-, Beta- and Gamma-Ray Spectroscopy
Alpha-, Beta- and Gamma-Ray Spectroscopy
Alpha-, Beta- and Gamma-Ray Spectroscopy
Ebook1,782 pages19 hours

Alpha-, Beta- and Gamma-Ray Spectroscopy

Rating: 1 out of 5 stars

1/5

()

Read preview

About this ebook

Alpha-, Beta- and Gamma-Ray Spectroscopy Volume 1 offers a comprehensive account of radioactivity and related low-energy phenomena. It summarizes progress in the field of alpha-, beta- and gamma-ray spectroscopy, including the discovery of the non-conservation of parity, as well as new experimental methods that elucidate the processes of weak interactions in general and beta-decay in particular. Comprised of 14 chapters, the book presents experimental methods and theoretical discussions and calculations to maintain the link between experiment and theory. It begins with a discussion of the interaction of electrons and alpha particles with matter. The book explains the elastic scattering of electrons by atomic nuclei and the interaction between gamma-radiation and matter. It then introduces topic on beta-ray spectrometer theory and design and crystal diffraction spectroscopy of nuclear gamma rays. Moreover, the book discusses the applications of the scintillation counter; proportional counting in gases; and the general processes and procedures used in determining disintegration schemes through a study of the beta- and gamma-rays emitted. In addition, it covers the nuclear shell model; collective nuclear motion and the unified model; and alpha-decay conservation laws. The emissions of gamma-radiation during charged particle bombardment and from fission fragments, as well as the neutron-capture radiation spectroscopy, are also explained. Experimentalists will find this book extremely useful.
LanguageEnglish
Release dateDec 2, 2012
ISBN9780444596994
Alpha-, Beta- and Gamma-Ray Spectroscopy

Related to Alpha-, Beta- and Gamma-Ray Spectroscopy

Related ebooks

Chemistry For You

View More

Related articles

Reviews for Alpha-, Beta- and Gamma-Ray Spectroscopy

Rating: 1 out of 5 stars
1/5

1 rating0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Alpha-, Beta- and Gamma-Ray Spectroscopy - K. Siegbahn

    ALPHA-, BETA- AND GAMMA-RAY SPECTROSCOPY

    VOLUME 1

    KAI SIEGBAHN

    Professor of Physics, University of Uppsala

    Table of Contents

    Cover image

    Title page

    Copyright

    PREFACE

    LIST OF CONTRIBUTORS

    INTRODUCTION

    Chapter 1: INTERACTION OF ELECTRONS AND α-PARTICLES WITH MATTER

    Publisher Summary

    INTERACTION OF ELECTRONS

    INTERACTION OF α-PARTICLES

    Chapter 2: INTERACTION OF γ-RADIATION WITH MATTER

    Publisher Summary

    § 1 Introduction

    § 2 Photoelectric effect

    § 3 Scattering by electrons

    § 4 Pair production

    § 5 Nuclear interactions and Delbrück scattering

    § 6 Total attenuation coefficients

    § 7 Absorption methods

    Chapter 3: BETA-RAY SPECTROMETER THEORY AND DESIGN. MAGNETIC ALPHA-RAY SPECTROSCOPY. HIGH RESOLUTION SPECTROSCOPY

    Publisher Summary

    § 1 Introduction. General discussion of spectrometers

    § 2 Some basic relations in β-spectroscopy

    § 3 Nomenclature

    § 4 The semicircular spectrometer

    § 5 The double focusing spectrometer

    § 6 Sector field spectrometers

    § 7 Some proposed high transmission instruments

    § 11 Precision spectroscopy

    Chapter 4: CRYSTAL DIFFRACTION SPECTROSCOPY OF NUCLEAR γ-RAYS

    Publisher Summary

    § 1 Introduction

    § 2 Theory of crystal diffraction spectrometers

    § 3 Curved crystal spectrometers

    § 4 A flat crystal spectrometer

    Chapter 5: THE SCINTILLATION METHOD

    Publisher Summary

    § 1 The scintillation counter

    § 2 Methods for charged particle spectrometry

    § 3 Methods for γ-rays

    § 4 Factors affecting resolution

    § 5 Crystal preparation and mounting

    § 6 Crystal efficiency and linearity

    § 7 Coincidence spectrometry

    CHAPTER VI: PARTICULAR DETECTION METHODS

    Chapter 6: PROPORTIONAL COUNTERS AND PULSE ION CHAMBERS

    Publisher Summary

    § 1 Introduction

    § 2 Volts per ion pair, W

    § 3 Statistical fluctuations in pulse size and resolution

    § 4 Ion chambers for alpha particle spectroscopy

    § 5 Proportional-counter alpha-spectroscopy

    § 6 Proportional counters for beta and gamma-ray spectroscopy

    § 7 Low background counting

    § 8 Carbon-14 age determinations

    § 9 Other applications of proportional counters

    § 10 High-pressure proportional counters

    § 11 High-temperature counters

    § 12. 4 π counters for absolute activity measurements

    § 13 Conclusion

    Chapter 7: SEMICONDUCTOR PARTICLE SPECTROMETERS

    Publisher Summary

    § 1 Introduction

    § 2 Charge collection

    § 3 Detector noise

    § 4 Radiation damage

    § 5 Electronics

    § 6 Detector types

    § 7 Application to nuclear physics

    CHAPTER VII: SOME EXPERIMENTAL TECHNIQUES

    Chapter 8: SAMPLE AND WINDOW TECHNIQUE

    Publisher Summary

    § 1 Introduction

    § 2 Thin foil and film technique

    § 3 Estimation of foil thickness

    § 4 Deposition technique

    § 5 Window absorption correction

    § 6 Acceleration of electrons

    Chapter 9: INTENSITY DETERMINATION OF PHOTOGRAPHICALLY RECORDED CONVERSION LINES

    Publisher Summary

    Chapter 10: MULTI-CHANNEL PULSE-AMPLITUDE ANALYZERS

    Publisher Summary

    § 1 Introduction

    § 2 Multi-discriminator analyzers

    § 3 Gray-wedge analyzers

    § 4 Amplitude-time conversion

    § 5 Computer storage techniques applied to analyzers

    § 6 Description of a typical analyzer with magnetic-core storage

    § 7 Multi-parameter analyzers

    § 8 Multi-sealer applications of analyzers

    Chapter 11: MEASUREMENT OF SOURCE STRENGTH

    Publisher Summary

    § 1 Introduction

    § 2 Direct counting methods

    § 3 Coincidence counting

    § 4 Choice of counting method

    § 5 Indirect methods

    CHAPTER VIII: PROCEDURES FOR THE INVESTIGATION OF DISINTEGRATION SCHEMES

    Chapter 12: GENERAL PROCEDURES

    Publisher Summary

    § 1 Introduction

    § 2 The preparation of the source material

    § 3 The calibration of instruments

    § 4 The search for γ-rays and the measurement of their energy

    § 5 The determination of the relative intensity of γ-rays

    § 6 The measurement of the β-rays

    § 7 The Fermi plot; end points and β-ray groups

    § 8 Determination of comparative half-lives ft

    § 9 Determination of internal conversion coefficients

    § 10 Examples of some disintegration schemes

    Chapter 13: SCINTILLATION SPECTRA ANALYSIS

    Publisher Summary

    § 1 Introduction

    § 2 General considerations

    § 3 Measurement of γ-ray scintillation spectra

    § 4 Determination of the response function

    § 5 Analysis of a spectrum by the peeling method

    § 6 Determination of energies and intensities

    7 Special applications

    § 8 Beta-rays

    § 9 Computer analysis of scintillation spectra

    Chapter 14: THE COINCIDENCE METHOD

    Publisher Summary

    § 1 Introduction

    § 2 Definitions

    § 3 Principles; errors

    § 4 Hardware

    § 5 A versatile coincidence circuit arrangement

    § 6 Geometries in electron–γ-ray coincidences

    § 7 Geometries in γ-γ coincidences

    § 8 Pile-up dangers

    § 9 Applications

    CHAPTER IX: THE SHELL MODEL

    Chapter 15: SHELL CLOSURE AND jj COUPLING

    Publisher Summary

    § 1 Introduction

    § 2 Magic numbers, or shell numbers

    § 3 Level sequence in the shell model

    § 4 Remarks on spin–orbit coupling

    § 5 Parity and orbital angular momentum

    § 6 Nuclei with incompletely filled shells

    § 7 Pairing energy

    § 8 Level occupation scheme

    § 9 Isomerism in nuclei of odd A

    § 10 Beta-decay

    § 11 Alpha-decay

    § 12 Center of mass motion

    Chapter 16: INTERMEDIATE COUPLING

    Publisher Summary

    § 1 Modification of the independent-particle model

    § 2 Intermediate coupling calculations

    § 3 Rotational interpretation

    § 4 Other interaction calculations

    Chapter 17: EVOLUTIONARY TRENDS IN NUCLEAR SPECTROSCOPY

    Publisher Summary

    Chapter 18: COLLECTIVE NUCLEAR MOTION AND THE UNIFIED MODEL

    Publisher Summary

    § 1 Introduction. General trends of nucleonic correlations

    § 2 Polarization effects of particles outside of closed shells

    § 3 Spherical nuclei

    § 4 Nuclei with stable quadrupole deformations

    § 5 Microscopic theory of collective motion and the theoretical estimates of the collective parameters

    Acknowledgement

    Chapter 19: ALPHA-DECAY

    Publisher Summary

    § 1 Introduction

    § 2 Alpha-decay conservation laws

    § 3 Decay energies and spectra of α-emission

    § 4 Treatment of α-decay data

    § 5 Decay rates of spherical nuclei

    § 6 Decay rate theory including non-central interactions

    § 7 Decay rates of odd-mass and odd–odd nuclei

    Chapter 20: GAMMA-RADIATION FROM CHARGED PARTICLE BOMBARDMENT; COULOMB EXCITATION

    Publisher Summary

    § 1 Introduction

    § 2 Coulomb excitation

    § 3 Gamma-radiation from nuclear reactions

    § 4 Gamma-ray energy and intensity measurements

    § 5 Particle-γ coincidences

    § 6 Angular distributions and angular correlations*

    § 7 Gamma-ray lifetime measurements

    Chapter 21: NEUTRON CAPTURE RADIATION SPECTROSCOPY

    Publisher Summary

    § 1 Introduction

    § 2 The neutron capture process

    § 3 Experimental arrangements for thermal γ-ray spectra

    § 4 Internal conversion and pair production

    § 5 Coincidence techniques

    § 6 Properties of levels

    § 7 Decay scheme of the compound nucleus

    Chapter 22: GAMMA-RADIATION FROM FISSION

    Publisher Summary

    § 1 Introduction

    § 2 Experiments and results

    § 3 Interpretation

    § 4 Discussion

    APPENDIX 1: GAMMA-RAY ATTENUATION COEFFICIENTS

    ATTENUATION COEFFICIENTS OF VARIOUS ELEMENTS

    ATTENUATION COEFFICIENTS OF VARIOUS SUBSTANCES

    THE EXPERIMENTAL RATIO (τa /τK)K edge OBTAINED FROM KIRCHNER

    APPENDIX 2: TABLES OF ELECTRON BINDING ENERGIES AND KINETIC ENERGY VERSUS MAGNETIC RIGIDITY

    ELECTRON BINDING ENERGIES

    KINETIC ENERGY VERSUS MAGNETIC RIGIDITY FOR ELECTRONS

    Copyright

    © North-Holland Publishing Company - 1968

    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 or otherwise, without the prior permission of the copyright owner.

    North-Holland ISBN: 0 7204 0083 x

    American Elsevier ISBN: 0 444 10695 2

    PUBLISHERS:

    NORTH-HOLLAND PUBLISHING COMPANY

    AMSTERDAM · NEW YORK · OXFORD

    DISTRIBUTORS FOR THE U.S.A. AND CANADA:

    ELSEVIER NORTH-HOLLAND INC.

    52 VANDERBILT AVENUE

    NEW YORK, N.Y. 10017

    1st edition 1965

    2nd printing 1966

    3rd printing 1968

    4th printing 1974

    5th printing 1979

    PRINTED AND BOUND IN GREAT BRITAIN

    BY W & J MACKAY LIMITED, CHATHAM

    PREFACE

    Ten years have passed since ‘Beta- and Gamma-Ray Spectroscopy’ was written. The present book includes α-spectroscopy and several other additional topics and thus presents a complete account of the whole field of radioactivity and related low energy phenomena. Actually, so much progress has been made during the past decade that this book has to be regarded as a new book rather than a revised edition of the old one. The discovery of the non-conservation of parity was, for instance, made after the publication of the old book which, in point of fact, served as the source of reference. New experimental methods, partly based on the consequences of the discovery, were devised and these contributed greatly to a better understanding of the processes of β-decay and weak interactions in general. The large field of nuclear spectroscopy as it concerns the lower excitation states reached in radioactivity and low energy reactions, has been considerably extended in various directions. Through developments in high resolution spectroscopy and by means of a number of improved or new methods, much more detailed knowledge of the structure of nuclear levels has been obtained. Successful models describing nuclear matter continue to be developed. It is hoped that this book on α-, β- and γ-ray spectroscopy will fully and correctly reproduce the present situation and serve as the starting point for new advances.

    Like the previous book the new one is primarily meant for experimentalists. Experimental methods and theoretical discussions and calculations are presented in such a way as to keep the link between experiment and theory unbroken. Generally the results, rather than detailed derivations, of theoretical calculations have been stressed.

    The large amount of material is divided into 26 chapters and 9 appendices, and is summarized in the introduction. The first volume is mainly concerned with general nuclear spectroscopy related to level schemes, models etc., whereas the second contains topics like transition rates, conversion, angular correlations, β-decay etc. The distribution of the chapters between the two volumes is not arbitrary, and it is hoped that a certain balance between the two volumes has been reached so that it should be easy to find where to look for the different topics. The appendices are placed in the volume to which they belong.

    As the editor, I have had the privilege of co-operating in this enterprise with a large number of friends and colleagues, some of whom were already active in the field in the early days before the shell theory. Most of my co-authors in the previous bookhave also contributed to the present one but the number of authors has increased from 42 to 77!

    The size of the book has grown in about the same proportion. The problem of keeping this large team together at an adequate speed of collective motion turned out to increase at least linearly with the number of individuals, with a considerable spread in individual motion. Some of the most diligent and the quickest in the team have no doubt been wondering what has happened in the queue behind. For all my co-workers and for myself it is therefore a source of satisfaction and a great relief to find that our common task is now complete. We sincerely hope that our readers will find the book useful.

    I am very much indebted to Dr R. M. Steffen and Dr C S. Wu who took part in the planning of Chapter XXIV, to Dr S. G. Nilsson (Lund) and Dr B. H. Nagel (Stockholm) who made valuable comments about the Introduction, and to Fil. Lic. S. E. Karlsson in my laboratory who was of considerable help in preparing Chapter III.

    To my friends at the North-Holland Publishing Company, in particular to its Director Mr M. D. Frank, and to his co-workers Mr W. H. Wimmers and Mr A. C. Pouwels, it is a real pleasure to extend a hearty thanks for their patience, encouragement and co-operation.

    KAI SIEGBAHN,     Uppsala

    March 1964

    LIST OF CONTRIBUTORS

    D. ALBURGER,     Brookhaven National Laboratory, Associated Universities, Inc., Upton, L.I., N.Y., U.S.A.

    R.A. ALLEN,     United Kingdom Atomic Energy Authority, Wantage Research Laboratory (A.E.R.E.), Wantage, Berkshire, England

    G. BÄCKSTRÖM,     Institute of Physics, University of Uppsala, Uppsala, Sweden

    I.M. BAND,     Physico-Technical Institute, Academy of Sciences, Leningrad, U.S.S.R.

    P.R. BELL,     Oak Ridge National Laboratory, P.O. Box Y, Oak Ridge, Tenn., U.S.A.

    R.E. BELL,     Radiation Laboratory, McGill University, Montreal 2, P.Q., Canada

    I. BERGSTRÖM,     Research Institute of Physics, Stockholm 50, Sweden

    S. BERKO,     Department of Physics, Brandeis University, Waltham 54, U.S.A.

    A. BOHR,     Institute for Theoretical Physics, Blegdamsvej 17, Copenhagen Ø, Denmark

    S.C. CURRAN,     The Royal College of Science and Technology, Glasgow, C.1., Scotland

    C.M. DAVISSON,     U.S. Naval Research Laboratory, Washington 25, D.C. 20390, U.S.A.

    M. DEUTSCH,     Physics Department, Massachusetts Institute of Technology, Cambridge 39, Mass., U.S.A.

    P.F. DONOVAN,     Bell Telephone Laboratories, Murray Hill, N.J., U.S.A.

    G.T. EWAN,     Atomic Energy of Canada Ltd., General Physics Branch, Chalk River, Ontario, Canada

    M. FERENTZ,     Physics Department, Jamaica 32, N.Y., U.S.A.

    H. FRAUENFELDER,     Department of Physics, University of Illinois, Urbana, Illinois, U.S.A.

    T.R. GERHOLM,     Institute of Physics, University of Stockholm, Stockholm Va, Sweden

    W.M. GIBSON,     Bell Telephone Laboratories, Murray Hill, N.J., U.S.A.

    R.K. GIRGIS,     Faculty of Science, University of Cairo, Giza, Egypt

    M. GOEPPERT MAYER,     Department of Physics, School of Science and Engineering, University of California, San Diego, La Jolla, Calif., U.S.A.

    M. GOLDHABER,     Brookhaven National Laboratory, Associated Universities, Inc., Upton, L.I., N.Y., U.S.A.

    F.S. GOULDING,     Lawrence Radiation Laboratory, University of California, Berkeley 4, Calif., U.S.A.

    R.L. GRAHAM,     Atomic Energy of Canada Ltd., General Physics Branch, Chalk River, Ontario, Canada

    L. GRODZINS,     Laboratory for Nuclear Science, Massachusetts Institute of Technology, Cambridge 39, Mass., U.S.A.

    S.R. DE GROOT,     Institute of Theoretical Physics, University of Amsterdam, Amsterdam, Holland

    S. HAGSTRÖM,     Institute of Physics, University of Uppsala, Uppsala, Sweden

    W.J. HUISKAMP,     Kamerlingh-Onnes Laboratory, Leiden, Holland

    J.H.D. JENSEN,     Institut für theoretische Physik, Philosophenweg 16, Heidelberg, Germany

    S.A.E. JOHANSSON,     Institute of Physics, University of Lund, Lund, Sweden

    E. KARLSSON,     Institute of Physics, University of Uppsala, Uppsala, Sweden

    P. KLEINHEINZ,     Institute of Physics, University of Uppsala, Uppsala, Sweden

    G. KNOP,     Physikalisches Institut der Universität Bonn, Bonn, Germany

    J.W. KNOWLES,     Atomic Energy of Canada Ltd., Neutron Physics Branch, Chalk River, Ontario, Canada

    O. KOFOED-HANSEN,     Department of Physics, Research Establishment Risö, Atomic Energy Commission, Roskilde, Denmark

    E.J. KONOPINSKI,     Department of Physics, Indiana University, Bloomington, Indiana, U.S.A.

    D. KURATH,     Argonne National Laboratory, Argonne, Illinois, U.S.A.

    R. VAN LIESHOUT,     Instituut voor Kernphysisch Onderzoek, Ooster Ringdijk 18, Amsterdam-O, Holland

    I. LINDGREN,     Institute of Physics, University of Uppsala, Uppsala, Sweden

    J. LINDSKOG ,     Institute of Physics, University of Uppsala, Uppsala, Sweden

    M.A. LISTENGARTEN ,     Physico-Technical Institute, Academy of Sciences, Leningrad, U.S.S.R.

    K.G. MALMFORS,     Research Institute of Physics, Stockholm 50, Sweden

    G.L. MILLER,     Bell Telephone Laboratories, Murray Hill, N.J., U.S.A.

    A.C.G. MITCHELL,     †, Department of Physics, Indiana University, Bloomington, Indiana, U.S.A.

    R.L. MÖSSBAUER,     California Institute of Technology, Pasadena, Calif., U.S.A.

    S. MOSZKOWSKI,     Department of Physics, University of California, Los Angeles 24, Calif., U.S.A.

    B.R. MOTTELSON,     Nordita, Copenhagen, Denmark

    H. MOTZ,     University of California, Los Alamos Scientific Laboratory, Box 1663, Los Alamos, New Mexico, U.S.A.

    O. NATHAN,     Institute for Theoretical Physics, Blegdamsvej 17, Copenhagen Ø, Denmark

    J.H. NEILER,     Oak Ridge Technical Enterprises Corporation, P.O. Box C, Oak Ridge, Tenn., U.S.A.

    W.A. NIERENBERG,     Department of Physics, University of California, Berkeley 4, Calif., U.S.A.

    S.G. NILSSON,     Institute for Theoretical Physics, University of Lund, Lund, Sweden

    C. NORDLING,     Institute of Physics, University of Uppsala, Uppsala, Sweden

    W.C. PARKER,     Institute of Physics, Chalmers Institute of Technology, Gothenburg, Sweden

    W. PAUL,     Physikalisches Institut der Universität Bonn, Nussallee 6, Germany

    B.G. PETTERSSON,     Institute of Physics, University of Stockholm, Stockholm, Sweden

    H. PRIMAKOFF,     Department of Physics, University of Pennsylvania, Philadelphia 4, Penn., U.S.A.

    J.O. RASMUSSEN,     Lawrence Radiation Laboratory, University of California, Berkeley 4, Calif., U.S.A.

    F. REINES,     Case Institute of Technology, University Circle, Cleveland 6, Ohio, U.S.A.

    R.A. RICCI,     Istituto Nazionale Energia Nucleare, Mostra d’Oltremare, Naples, Italy

    M.E. ROSE,     University of Virginia, McCormic Road, Charlottesville, Virginia, U.S.A.

    S.P. ROSEN,     Department of Physics, Purdue University, Lafayette, Indiana 47907, U.S.A.

    N. ROSENZWEIG,     Argonne National Laboratory, Argonne, Illinois, U.S.A.

    K. SIEGBAHN,     Institute of Physics, University of Uppsala, Uppsala, Sweden

    H. SLÄTIS,     Research Institute of Physics, Stockholm 50, Sweden

    L.A. SLIV,     Physico-Technical Institute, Academy of Sciences, Leningrad, U.S.S.R.

    A.H. SNELL,     Oak Ridge National Laboratory, P.O. Box Y, Oak Ridge, Tenn., U.S.A.

    P. SPARRMAN,     Institute of Physics, University of Uppsala, Uppsala, Sweden

    R.M. STEFFEN,     Physics Department, Purdue University, Lafayette, Indiana, U.S.A.

    T. SUNDSTRÖM,     Institute of Physics, University of Uppsala, Uppsala, Sweden

    A.W. SUNYAR,     Brookhaven National Laboratory, Associated Universities, Inc., Upton, L.I., N.Y., U.S.A.

    J.E. THUN,     Institute of Physics, University of Uppsala, Uppsala, Sweden

    H.A. TOLHOEK,     Natuurkundig Laboratorium der Rijks-Universiteit te Groningen, Westersingel 34, Groningen, Holland

    A.H. WAPSTRA,     Instituut voor Kernphysisch Onderzoek, Ooster Ringdijk 18, Amsterdam-O, Holland

    H.W. WILSON,     Scottish Research Reactor Centre, East Kilbride, near Glasgow, Scotland

    R. WILSON,     Harvard University, Cyclotron Laboratory, Cambridge 38, Mass., U.S.A.

    C.S. WU,     Department of Physics, Columbia University, 538 West 120th Street, New York 27, N.Y., U.S.A.

    M.J.L. YATES,     A.W.R.E., Aldermaston, Berks., England

    INTRODUCTION

    KAI SIEGBAHN

    Our knowledge concerning nuclear structure and nuclear forces comes from three main sources: (1) radioactive decay, (2) nuclear reactions and (3) high energy or elementary particle reactions. There is a certain overlap, particularly with the first two of these, both with regard to the experimental techniques and to the type of information which is obtained. The third one is gradually drifting away from the other two and becoming a study of a new physical domain, in some respects equally distant from low energy nuclear physics as this is from atomic shell physics.

    Alpha-, beta- and gamma-ray spectroscopy is often used synonymously to the first mentioned field, although these radiations, of course, also occur in nuclear reactions. In radioactive decay only the lowest nuclear levels are excited, up to some few MeV. It is important to bear this in mind because of the inherent limitations to the information which result from it. The phenomena which are accessible for study at these energies are, however, of basic importance for the understanding of nuclear structure in general. Furthermore, they are not masked by the increasing complexity occurring at higher excitations with gradual deterioration into statistical processes.

    The concentration on the purely low energy phenomena occurring in radioactivity has actually been of great value for the successful development of nuclear theories during recent years. One may recall the confused situation at the time of the break-through of the nuclear shell theory at the end of the forties. It was then suddenly realized that one of the most fundamental aspects of nuclear matter had been hidden under the heavy bombardment of nuclear artillery.

    It would, however, not be justified to overstress this point. At present, one is fairly well aware of the applicabilities and natural limits of the concepts describing nuclear behaviour at different energies. In view of this, one of the present trends is to extend low energy measurements, such as partly developed in radioactivity work, to the region of higher excitation energies accessible by nuclear reactions. A more complete picture of the excitation pattern can then be obtained. Not only can the higher excitation stages be reached but many new lower stages will also show up which have not been populated in radioactive decays due to the operation of selection rules.

    A large variety of techniques, experiments and theories has developed out of the field of α-, β- and γ-ray spectroscopy. A fairly disperse group of phenomena has been subjected to studies. Nuclear structure, and the quantum mechanical properties of the nuclear ground states and excited levels, constitute the most important and also by far the largest field. Continuous progress in spectroscopic resolution now allows a detailed insight into the level schemes. There has been a most successful period of interesting developments of nuclear structure models.

    The universal interaction in β- and muon decay and muon capture is a phenomenon of fundamental significance. One of the most dramatic developments has been in the field of weak interactions centering around the discovery of non-conservation of parity. The conclusions drawn from these studies have had far-reaching consequences resulting in the formulation of basic symmetry laws of nature.

    The recent experimental discoveries of two types of neutrinos, one associated with the electron and the other with the muon, will naturally be of decisive importance for the future development of the theory of weak interactions.

    The absorption and scattering of nuclear radiation in matter leads to the study of phenomena which branch out in many directions. One of the most spectacular discoveries in recent years is the nuclear resonance scattering of γ-rays without nuclear recoil, the Mössbauer effect, with wide applications in the theory of relativity, solid state physics and chemistry.

    These are some of the most interesting lines of development during recent years and the list could be extended with numerous other examples: the detailed description of the interaction between the nucleus and the atomic shell, transition probabilities, higher order quantum electrodynamic processes, positronium, nuclear recoils, electromagnetic radiation emitted during particle and neutron bombardment and during fission, angular correlations perturbed by internal and external fields, nuclear alignments, and the developments of precision methods and new detection techniques to deal with all these problems. The aim of the present book is to give a fairly complete account of the situation at the various fronts. A short summary is presented here in order to give a condensed picture of the numerous topics treated in the book.

    Nuclear spectroscopy can be traced back to about fifty years ago. Though the existence of α-, β- and γ-radiation emitted from radioactive elements became known soon after the discovery of radioactivity, spectroscopic work did not begin until 1911. It was found by various methods that β-radiation contains both a continuous component and discrete lines. The continuous spectrum was for a long time completely unexplained, whereas at the beginning of the 1920’s the line spectrum was found to be due to an interaction of the atomic shell with the decaying nucleus. The existence of long ‘series’ of β-lines, the energy differences of which corresponded to differences in the atomic shells of the daughter product, led to an interpretation of the β-lines as due to photoelectrons expelled by monochromatic γ-quanta emitted from the nucleus. Later on, when the process of internal electron conversion was theoretically studied, it was found that this simple description did not hold. Conversion must instead be regarded as an additional mode of nuclear deexcitation, i.e. the nucleus can be de-excited either by the emission of a γ-quantum or by a direct interaction with the extranuclear electrons, in which case a conversion electron from one of the atomic shells is emitted. This view was more directly verified when it was experimentally found that the radioactive decay constant of technetium could be changed by varying the wave function of the atomic electrons at the nucleus by chemical means or by means of a very high pressure. A two-step photoelectric process would have left the decay rate unaffected by such changes of the external conditions.

    The conversion phenomenon has been found to be extremely useful in the study of the finer details of radioactive decay. Not only do the measurements of the energies of the β-lines permit direct calculation of the corresponding energies of the nuclear transitions, which yield fundamental information on the decay scheme, but a closer study of the intensities of the lines also allows determination of the quantum mechanical properties of the different nuclear states, such as spin and parity.

    In 1934 the first acceptable theory for the continuous spectrum was presented. It was based on the neutrino hypothesis and was quite satisfactorily able to account for the relation between the decay rate and maximum energy of the β-particles. Furthermore, the shape of the spectrum was calculated. The theory was soon extended to all possible types of interactions between heavy and light particles in β-decay. The theory contained, however, several ambiguities, which could only be removed by experiment. On the other hand, the experimental methods at that time were not developed to such an extent that fruitful work in nuclear spectroscopy was possible. It is true that some interesting information had already been obtained from the conversion spectra of natural radioactive species and from the fine structure of α-spectra discovered in 1929, but the conclusions that could be drawn from these measurements did not give any deeper insight into the decay mechanism.

    When artificially produced isotopes became generally available during the forties, the domain of nuclear spectroscopy developed greatly. To start with, the lack of adequate experimental methods and the difficulties of handling such sources resulted in numerous erroneous conclusions concerning conversion- and γ-lines and other features of the decays. Great efforts have therefore continuously been made to develop instruments with which experimental data of high quality can be obtained. High resolution magnetic spectrometers have been designed, which, together with improved coincidence methods and more advanced radiochemical procedures and isotope separations, yield conclusive data on nuclear disintegration schemes. When the conversion rates are too low to be detected, the energies of the different nuclear transitions can be determined accurately from energy measurements on secondary electrons, which the γ-rays expel from a thin converter in the spectrometer. The crystal diffraction method, previously used for X-ray studies, has been successfully modified to meet the difficulties encountered at short wavelengths. Unconverted γ-radiation can be studied at a good resolution by this method, provided that very strong sources are available. Refined proportional counter techniques and scintillation counter methods, introduced at the end of the forties, greatly extended the experimental possibilities. The high efficiencies of these counters in counting γ-quanta and the rapid responses make them extremely useful in many sorts of coincidence experiments. Further, the energy distribution of the radiation can be studied by pulse height analyzers. Because of the statistical nature of the registration, the resolution is much less than with magnetic spectrometers, being about 10% at 0.5 MeV and then decreasing as the square root of the energy. Magnetic spectrometers, on the other hand, require much stronger sources, in particular when low-converted γ-rays are investigated. Under favourable conditions, it has been found possible to develop the resolving power of special devices to the limit set by the inherent width of atomic levels, i.e. relative line widths of ≈ 1:10⁴. Already at a somewhat lower resolution, conversion electron spectra of great complexity can be disentangled by means of magnetic spectrometers. At higher energies (> 1 MeV) one can use the pair process for studying the γ-radiation, either in conjunction with scintillation counters or with magnetic spectrometers. Convenient coincidence arrangements for the purpose of distinguishing other processes have, in both cases, been developed.

    It has been found that most problems in nuclear spectroscopy cannot be successfully tackled by one single instrument or one particular method. Instead, the nuclear spectroscopist must view the problem from different angles and judge the techniques or even combination of methods to be used. As an example one can mention that very efficient ways of combining magnetic spectrometers with scintillation counters as detectors and coupled in coincidence arrangements have been used. In this way one can couple the good resolution of a magnetic spectrometer with the high efficiency or the fast response of a scintillation counter.

    A new powerful development is the semiconductor detector. Since the energy required to create an ion pair in such a device is much less than in a scintillation crystal, the attainable resolution can be considerably increased. The efficiency is less and the response slower, however. The amplifier noise is one of the limiting factors. Resolution better than 5 keV (being approximately independent of energy) has been reported under special conditions.

    Many-parameter-multichannel analyzers have also recently been developed which will make many new types of experiments feasible.

    By the end of the forties a certain amount of experimental data on disintegration schemes had been accumulated. This material was very incomplete and in many cases of low accuracy. Spin and parity assignments of many well-established nuclear levels had been made, in particular by the application of the selection rules for β- and γ-transitions (using ft-values and γ-ray intensities) and by means of conversion data. The theoretical background for such considerations was at that time uncertainly based. The theory of transition probabilities of multipole radiation was not very advanced and the calculated conversion coefficients were approximate. The most serious difficulty in nuclear spectroscopy at that time, however, was the lack of any useful nuclear model. No really fruitful systematic work was therefore possible and there was only a loose connection between the results of experimental investigations and nuclear structure. This picture changed completely with the advancement of the nuclear shell theory, later complemented with the concept of collective nuclear motion. In order to explain the existence of the magic numbers of neutrons and protons in nuclei, it was assumed that an odd nucleon could move independently in the nucleus in a potential determined by the average field of the remainder. By further assuming a strong spin–orbit coupling, it was found possible to find the correct sequence of quantum states when the nucleons were successively added according to the Pauli principle. This model met with considerable success in explaining, for example, the spins of odd mass nuclei.

    In the original shell theory, one further fact was taken into consideration, namely the coupling of the nucleons two and two to angular momentum zero. This kind of pairing essentially amounts to consideration of the ‘diagonal’ effect of the pairing interaction, while the more recent pair correlation theory also takes the correlation between the pairs into account, distributing them over several different configurations. The pairing within a given configuration was, in the earlier stage of the theory, a reflexion of the well-known empirical fact that nuclei containing paired protons or neutrons are energetically favoured compared to even–odd or odd–odd nuclei. The ‘pairing energy’ is thus a binding energy, of about 1–3 MeV, and increases in the shell theory with angular momentum j.

    Even at an early stage, attempts were made to include residual forces of a realistic type, that is of finite range and containing exchange mixtures. A complete treatment of such a residual force is not limited solely to the inclusion of the diagonal, but also includes non-diagonal elements and therefore leads to ‘configurational mixing’. An exact treatment of this problem is of a complexity that increases with the number of particles. Several calculations, in particular on light nuclei, have been made with good success. For heavy nuclei, calculations have been made on nuclei in the neighbourhood of closed shells, in particular in the region around the ‘double’ magic Pb²⁰⁸ (82 protons and 126 neutrons) where good agreement with experiments has been found.

    The nucleon–nucleon interaction results in a splitting of the degenerate shell-model configuration, which can be of the same order of magnitude as the spin–orbit coupling. In regions far from the magic numbers in particular, several possible one-particle states start to compete, leading to space correlations between the nucleons. Instead of a spherically symmetric potential, a spheroidal potential is then a better approximation in these regions. Using such a potential, where the eccentricity enters as a parameter, a more realistic single-particle level scheme can be calculated. It is found that the removal of the degeneracy leads to significant changes in the level sequence at larger deformations. These theoretical level schemes are usually the starting point for the analysis of spectra characterized by single-particle excitation.

    The main effect of the very complicated realistic residual force can be schematically represented by the assumption of two very simple forces of definite ‘specificity’. One is the so-called quadrupole force accounting for the most important coherent long-range property of the nuclear force, apart from the monopole component included in the spherical potential. On the other hand, the pair force is characterized by a very short range and simulates the short-range characteristics of the residual nuclear force. If this effect dominates, the nucleus will have spherical symmetry. Alternatively, if the quadrupole forces dominate, which implies that essentially all the individual wave functions overlap in contrast to the pairwise overlap, the characterization of the pair coupling scheme, the different nucleons are concentrated in certain directions, that is, one has a deformed nucleus. The pairing effect depends linearly on the number of nucleon pairs, whereas the quadrupole effect, which is the combined result of all the nucleons’ mutual interactions, is quadratically dependent on the nucleons. Thus, while the pairing force is dominant for a few nucleons outside of closed shells, the quadrupole forces ultimately dominate when the number of nucleons outside of closed shells increases. The deformation is consequently largest far from closed shells (nucleons and nucleon holes are equivalent).

    The dominance of the nuclear quadrupole force ultimately leads to the realization of the deformed coupling scheme. At this limit, the effects of the long-range quadrupole forces are adequately described by a deformed nuclear field. The latter should be regarded as a convenient description rather than a picture in contradistinction to the configurational-mixing description. In a few cases it has also been possible to obtain nearly the same results for a particular problem by treating it from the ‘deformed’ side, describing the nuclear wave functions on the basis of a deformed field on the one hand, and from the spherical side using detailed residual forces and configurational mixing on the other. The best example of this is probably F¹⁹.

    The competition between the pairing coupling scheme and the quadrupole coupling scheme, as the number of extra-core nucleons is increased, is primarily responsible for the fact that the nucleus becomes soft with regard to changes in shape. This softness shows up in terms of low-lying vibrations about the spherical equilibrium configuration. The quantized quadrupole vibration has its first excited level with spin 2 and even parity. According to the harmonic description of the vibration, three close-lying levels should occur, characterized by the quantum states 0+, 2+, and 4+, corresponding to two-phonon excitations. Vibrations about non-spherical equilibrium shapes are also possible and have been classified into different types, one associated with vibrations that preserve rotational symmetry around the nuclear axis, and a further type of vibration that is associated with fluctuations about this cylindrically symmetric shape.

    A deformed nucleus which is excited can describe a quantized rotation around an axis perpendicular to the symmetry axis of the nucleus. Since the mass transport associated with this collective rotation is very large, the frequencies corresponding to the lowest angular momentum states are small, and the associated energies low compared to individual-particle excitation energies. In the regions of deformed nuclei, the level schemes consist of bands of rotational levels superimposed on single-particle excited levels. Evidently, the whole picture resembles that of molecular spectroscopy. The unification of the single-particle aspects and the collective aspects into one model is well described by the term ‘unified model’.

    The nuclear state is described in terms of the quantum numbers K, I and M. Since the potential is not spherically symmetric, the angular momentum j of the particle moving around the deformed core is not conserved. Its projection on the axis of symmetry is conserved, however, and is denoted K. The rotation contributes an angular momentum R perpendicular to the symmetry axis. K and R combine to give the total angular momentum I, the projection of which along any arbitrary axis is denoted by M.

    Many of the properties of strongly deformed nuclei can be formulated by simple expressions. The relative spacing between successive rotational levels can be described accurately by means of the quantum numbers I and K, the formulae showing close similarity to those known from molecular spectra. The energies of the states also include the moment of inertia. Experiments show that this moment is less than the rigid body value but greater than the hydrodynamic (irrotational flow) value. The relative quadrupole transition probabilities in rotational bands are also remarkably well reproduced by the theory. In order to describe the properties of the states such as magnetic moments and electric quadrupole moments, one needs, apart from the moment of inertia, the intrinsic quadrupole moment Q0, and the g-factors gK and gR describing the intrinsic and rotational magnetism.

    A new powerful formalism for the description of nuclear properties was introduced in 1958. It is based on the similarities between nuclear matter and matter in the superconductivity state. The well-known BCS theory, describing superconductivity, was primilarily adopted to account for the fact that there is a large energy gap in even–even nuclei between the ground state and the first excited state. This is analogous to the well-known energy gap found in superconductivity. Breaking pairing bonds between nucleons means creating quasi-particle states with the same quantum numbers as single-particle intrinsic states. The probabilities of such an empty single-particle state and a filled one are represented by the parameters U and V. This formalism offers a description of nuclear structure by means of which the different models and the residual interactions can be understood in a natural way. Some previous difficulties, such as the deviation of the moments of inertia, can be treated more satisfactorily. The calculated collective gyromagnetic ratios are greatly improved, and the variation in ft-values for transitions involving the same single-particle states as a function of mass number is better reproduced.

    The transition probabilities in β-decay are governed by selection rules and can be determined experimentally by measuring the partial half-lives of the different branches of the decay. This can be accomplished by means of the so-called Fermi analysis. One can then calculate the ft-values of the transitions, t being the partial half-life and f a function of the transition energy and the atomic charge. The ft-value is proportional to the inverse square of the nuclear matrix element. The ft-values of various β-transitions can be grouped into different classes. Using this classification and the particular selection rules which apply to the different classes, one can draw conclusions regarding the changes in spin, parity and orbital angular momentum associated with the transitions.

    As mentioned above, a fundamental problem with special bearing upon the nature of nuclear forces is the form of the Hamiltonian describing the interaction between heavy and light particles in β-decay. This problem was previously mostly studied by the shapes of the continuous spectra of forbidden transitions.

    The form of the interaction term in β-decay can also be studied by means of the angular correlation between electron and neutrino. In this case also, allowed transitions in principle yield the desired information. The low energies of the recoiling atoms (of the order of 100 eV) strongly favour the use of radioactive gases in such experiments. Only a few have been performed, but they show, among other things, that a continuous neutrino spectrum is emitted during β-decay, whereas monoenergetic neutrinos are emitted in K-capture.

    There are five possible types of relativistically invariant interactions in β-decay, described by a scalar (S), vector (V), tensor (T), axial vector (A) and a pseudoscalar (P), respectively. Experiments must determine the correct combination of the different invariants.

    The possible invariants do not all yield the same selection rules. Thus the scalar and vector interactions follow the so-called Fermi selection rules, stating that, for allowed transitions, the spin change is zero and the parity is unchanged. With the tensor and axial vector interactions the Gamow–Teller selection rules apply, in which case the spin change is either zero or one for an allowed transition, the parity again being unchanged. There is definite experimental evidence for both types of interactions. By comparing, for example, the empirically found ft-values of superallowed β-transitions occurring between mirror nuclei having closed shells plus or minus one nucleon, for which the matrix elements can be well estimated by means of single-particle wave functions, it can also be concluded that the Fermi and Gamow–Teller couplings give about equal contributions. Other evidence supports this result.

    Some years ago experimental evidence seemed to favour a combination of a scalar and a tensor interaction in β-decay. A new neutrino recoil experiment, however, gave a conflicting result and the confused situation was carefully reexamined. It was then found that the combination V–A was the only one which could fit the results of all the new experiments. Final and decisive evidence was delivered by an ingenious nuclear resonance scattering experiment performed on circularly polarized γ-radiation from Eu¹⁵²m. This experiment formed one of the high-lights of the ‘parity era’ starting with the discovery in 1957 that parity, contrary to all previous concepts, was not conserved in weak interactions such as β-decay. This was proved by the famous experiment on aligned Co⁶⁰. It was found that the electrons were preferentially emitted in one direction, namely opposite to the nuclear spin. Since nuclear spin is an axial vector and the linear momentum of an electron is a polar vector, the parity operation, i.e. a space reflexion, changes the system. There are several consequences of this phenomenon. Decay electrons from an unoriented source are emitted longitudinally polarized: negative electrons with ‘left-handed’ spins (negative ‘helicity’), i.e. opposite to its motion, and positrons with right-handed spins (positive helicity). The nuclear resonance scattering experiment on Eu¹⁵²m showed, incidentally, that the neutrino is a left-handed particle and that the antineutrino is a right-handed one. These circumstances form the basis for the so-called two-component neutrino theory. The polarization of electrons can be found by turning the spin axis through 90° in an external field with a subsequent scattering, showing left–right asymmetry. Another possibility is to observe the circular polarization of the bremsstrahlung produced when the electrons are absorbed.

    A large amount of experimental and theoretical work has been devoted to a deeper understanding of the parity law in weak interactions. Experiments also show that charge conjugation (particle→antiparticle) invariance is violated in these interactions, whereas there are indications that the combination of charge conjugation C and space reflexion P gives a conserved quantity. A redefinition of the parity law would then be that the laws of nature are invariant under the combined CP operation. This is equivalent to invariance under time reversal T as follows from a theoretical result known as the CPT theorem. This theorem states that, under some general conditions, a relativistically invariant theory is also invariant under the combined CPT operation.

    Muon decay and muon capture are of great interest in connection with the interaction in β-decay. A comparative study of these processes provides evidence for and against the validity of the postulate that the interaction between the four Fermi particles involved in the β-decay is a universal one, applying to any four fermions. The evidence seems at present to indicate such a universality. The finer details of the β-theory are still subjects for study. To explain the equal strengths of the polar vector interaction in β- and muon decays, it has been suggested that the polar vector current responsible for these weak interactions should be conserved, the so-called ‘conserved vector current’ theory (CVC). This theory has recently been supported by a careful examination of the shapes and βγ angular correlations (circularly polarized) of selected decays.

    The effect of the electronic shell on the decaying nucleus manifests itself in many ways, the two most obvious ones being electron capture and the conversion phenomenon. Since the former varies approximately as the inverse third power of the principal quantum number, one usually has to consider only K-capture. It competes with the process of positron emission and (except for large decay energies) is favoured energetically. This follows from the fact that an electron is captured instead of emitted as in positron emission, which makes electron capture the only possible mode of decay for sufficiently low decay energies. This energy is carried away by a neutrino, the observable effect being the emission of characteristic X-radiation or, alternatively, Auger electrons following the refilling of the atomic shell. In some cases it is possible to measure the decay energy by determining the upper energy limit of the accompanying weak, continuous, internal bremsstrahlung spectrum. The ratio of K-capture to positron emission can be determined from the intensity ratio of the X-rays or Auger electrons and positrons or by means of special coincidence experiments. For a given maximum positron energy, this ratio increases with Z and at a given Z it increases with decreasing maximum positron energy. The fact that the transition probability of electron capture depends on the electron density at the nucleus, just as in the case of internal conversion, has made it possible to change by chemical means the decay constant of another radioactive decay apart from the above mentioned Tc, namely in the case of Be⁷.

    A number of interesting high order effects may occur in radioactive decay. Normally, the energy in β-decay is shared between the β-particle and a neutrino. One can, however, visualize several competing processes of higher order. One is the above mentioned internal bremsstrahlung in which the β-particle is retarded in the coulomb field of the nucleus, resulting in the simultaneous emission of a bremsstrahlung quantum, a β-particle, and a neutrino. A process of an even higher order would be the emission of two such quanta. The internal compton effect is another high order effect. Here, a γ-quantum undergoes internal scattering (predominantly in the K-shell), resulting in the simultaneous emission of a γ-quantum of lower energy and an electron. Other high order effects may occur due to the creation of a hole in the K-shell after K-capture, or internal conversion immediately followed by the emission of another conversion electron or an electron pair. In the latter case, the electron in the pair has a certain small probability of being captured in the hole and the energy will then be carried away by the positron alone. This process would result in a positron line. Other possible high order effects are, for example, two-photon emission or two-conversion-electron emission. Some of the above processes, such as the internal bremsstrahlung, are well-known phenomena whereas some others are extremely difficult to observe or have not yet been established experimentally.

    Internal conversion is closely related to the process of γ-ray emission. Whereas the transition probabilities in the latter case include inaccurately known nuclear matrix elements, these cancel in the calculations of internal conversion coefficients, as the conversion coefficient is defined as a ratio of the conversion electron intensity to the γ-ray intensity. The radiating nucleus can be regarded as an oscillating electric and magnetic multipole. The contributions of the different multipoles to the radiation field are governed by selection rules. These concern the angular momentum of the photon multipole, its parity and, finally, the intrinsic properties of the nuclear wave function. The transition probability decreases rapidly for higher multipoles. Electric quadrupole or magnetic dipole transition rates correspond to extremely short emission times for spectroscopically typical transition energies (< 10–9 sec). Such experiments have been performed by means of the delayed coincidence technique in the nanosecond region using scintillation detectors, and with the nuclear resonance scattering method. The lifetimes are much longer for higher multipole transitions and are often classified as due to nuclear isomeric states. One of the first startling results of the nuclear shell theory was its ability to explain the so-called ‘isomeric islands’ on the isotope chart. Due to the spin–orbit coupling all subshells having j=l+½ are depressed relative to those with j=l−½. The level corresponding to the highest j. are of the M4 type, in accordance with experiment.

    , is a consequence of configurational mixing in this scheme and the measured matrix elements show a wide spread.

    In the pairing picture, the difference between the electric and magnetic multipole transitions is a consequence of the diffuseness factor (containing U and V) which differs in the Eλ and Mλ cases. Generally one might say that the estimate of the nuclear matrix element in terms of single-particle expressions is made extremely uncertain by the important role played by other residual forces than pairing.

    The enhanced transition probability due to cooperative nuclear motion in nuclei, for example, the rotational levels, manifests itself in a considerably reduced lifetime of the states compared to simple one-particle model calculations. The collective model can effectively account for these transition rates.

    It is extremely important for nuclear theory to have a knowledge of spin and parity changes in nuclear decay. Several methods exist, but the most useful one is to measure the internal conversion coefficients. These have been calculated for different multipole orders, energies and Z-values, and the fact that no nuclear matrix elements enter makes the numbers independent of any particular nuclear model. The K-shell conversion dominates for the lowest multipole orders. Experimentally, it is frequently easier to measure the ratio of the conversions in different shells (e.g. the K:L ratio) rather than the conversion rates themselves. If the conversion lines in the three subshells LI, LII and LIII can be resolved, one obtains detailed information regarding the radiation properties, such as the amount of admixture of electric and magnetic components.

    In some special cases when the usual radiation matrix elements are hindered, due to selection rules which do not apply to the matrix elements responsible for the interaction between the electron and the nuclear transition charges and currents inside the nucleus, the conversion may be nuclear structure dependent. Such ‘penetration’ effects have been found in several cases, in particular in El transitions among heavy elements.

    When the conversion rates are low and difficult to determine, directional correlation studies between successively emitted γ-quanta are particularly informative in assigning spin values. When polarization correlation studies are possible, information concerning parity changes is also obtained. Directional correlation studies involving conversion electrons yield both quantities. Beta–gamma directional correlations give information about the interaction in β-decay.

    For α-radiation, angular correlation experiments yield α-partial wave mixtures and phases. Thus, there is a large interference term in the αγ correlation due to admixture of the orbital angular momentum in the main group L=0 and L=2 as shown by experiments on the transuranic elements. This admixture is also suggested from low temperature alignments. These waves constructively interfere in the polar regions of the prolate spheroidal surface, showing a preference for α-emission through the thinner barrier at these places.

    Investigations of angular correlations have also led to other very interesting possibilities. The correlation may be perturbed, owing to electric and magnetic fields acting on the electric and magnetic moments of the nuclei. These fields may, at least in principle, either be calculated or, in the case of a magnetic field, externally applied. By these means it has become possible to measure magnetic moments and electric quadrupole coupling energies of excited nuclear states. In some cases, it has been possible to determine nuclear moments from the splitting of Mössbauer resonance lines. The ratio of the excited state and ground state magnetic moments is measured with reasonable accuracy if high internal fields are available. Electric quadrupole splittings can be determined, but require a knowledge of crystal wave functions for the evaluation of quadrupole moments. The precession of Mössbauer scattered γ-radiation can also be used for g-factor measurements. (See Perturbed Angular Correlations, edited by E. Karlsson, E. Matthias and K. Siegbahn; North-Holland Publishing Company, Amsterdam, 1964.) Magnetic moments of excited states have also been measured by aligning the nuclei at low temperatures.

    The spins and moments of the ground states of radioactive nuclei can be measured by atomic beam resonance techniques and by some other less generally applicable methods.

    It is of increasing importance to have accurate experimental determinations of the electric and magnetic moments of the excited and ground states of nuclei. For the excited states, unfortunately, no resonance method as yet exists and the accuracy attainable is consequently fairly modest. The number of levels which is accessible for study is also rather limited at present. This situation will probably change considerably when new methods for producing very strong (external or internal) magnetic fields are developed.

    Most of the experimental and theoretical approaches described in this book will no doubt be improved and extended in various directions in the future. For example, it is probable that an increasing amount of work will be devoted to the study of nuclear states of higher excitation energies. Our knowledge of nuclear matter, as deduced from the lowest excitation states, could be considerably enlarged by studying γ-radiation and conversion electron spectra emitted during particle bombardment. Ordinary coulomb excitation has already given much basic information on rotational quantum states. Heavy-ion bombardment is a powerful tool for the excitation of very high rotational states which could be explored more extensively.

    The possibility of the existence of an upper limit for the rotational bands should be investigated and also the possibility of a sudden or gradual increase in the moment of inertia as the frequency of rotation increases. Theoretically, one expects that, for high rotational frequencies, the Coriolis force will ultimately destroy the superfluid state of the nucleus, just as the increasing external magnetic field will finally destroy the superconductor state when it becomes sufficiently large. The region above the energy gap provides an interesting field of study. One can here investigate the effect of the breaking of nucleon pairs on the superfluidity, as exhibited in the associated moment of inertia values. This effect from the breaking of pairs corresponds to the temperature effect in the superconductor analogue.

    The information which can be obtained from studies of the decays of radioactive isotopes is still far from explored. Higher accuracy in determining the properties of the different nuclear states and also an extension of the various characteristics, such as electric and magnetic moments, transition rates etc., describing each level will presumably, together with the above discussed study of higher excitation states, lead to an improved understanding of nuclear matter than that we have to-day.

    I

    INTERACTION OF ELECTRONS AND α-PARTICLES WITH MATTER

    G. KNOP

    W. PAUL

    Publisher Summary

    The elastic scattering of electrons passing through matter can be divided roughly into four classes: (1) single scattering, (2) plural scattering, (3) multiple scattering, and (4) diffusion. When electrons of a definite energy pass through a foil of matter, their energy is decreased. In a study dsercibed in the chapter, the interaction of the incident electrons with the atomic electrons in the foil is characterized by the fact that the energy transferred to the atoms per collision is very small. In addition to the energy loss resulting from the excitation and ionization processes, the energy loss occurs due to the emission of bremsstrahlung also, which occurs when the electron is accelerated in the Coulomb field of a nucleus. In this chapter, the average energy loss per cm of path is given by Bethe’s formula. There are two important types of measurement—namely, the specific ionization and the total ionization.

    INTERACTION OF ELECTRONS

    Electrons penetrating matter lose energy and are deflected from their original course: they are scattered. Changes take place also in the matter that is penetrated, the constituent atoms are excited or ionized, and dissociation of molecules, changes in the lattice structure of crystals, changes in the conductivity, and many other secondary processes have been observed. These phenomena will be discussed here only if they give direct information about the energy of the electrons, as e.g. the ionization. Furthermore, we will restrict the discussion to electron energies in the range 10⁴–10⁷ eV, i.e. to the region of the radioactive β-emitters. For these energies, the deflexion of the electrons is due almost entirely to the elastic collisions with the atomic nuclei, while the energy loss, except that due to the bremsstrahlung, which is practically negligible, results from the interaction with the atomic electrons. Therefore it is possible to treat the two phenomena separately, though of course they always occur together. For positrons, the general behaviour is the same as for electrons. However, there are deviations which are mentioned at the corresponding places. A detailed review has been given by Bothe¹ and by Bethe and Ashkin², while the theoretical principles have also been discussed in detail by Sauter³. For general discussion of the theory see Mott and Massey⁴.

    § 1 Elastic scattering of electrons by atomic nuclei

    The elastic scattering of electrons passing through matter can be divided roughly into four classes: (1) Single scattering; (2) Plural scattering; (3) Multiple scattering; (4) Diffusion. If the thickness d of the layer is very small, d 1/σN, where σ is the cross-section and N the number of scattering atoms per cm³, we have practically only single scattering, i.e. nearly all the scattered electrons are scattered by only one nucleus. It should be remembered, however, that the cross-section for the scattering of electrons by nuclei decreases very strongly with increasing scattering angle, so that the relation given above for the thickness of the layer shows a pronounced dependence on the angle of scattering Θ. For larger values of the thickness, d ~ 1/σN, we get plural scattering, i.e. the probability that a given scattering angle is due to a number of successive single scattering processes becomes appreciable. When the thickness becomes so large that the mean number n of scattering processes becomes larger than about 20, we speak of multiple scattering. The angular distribution W(Θ) of the scattered electrons is approximately gaussian as long as the mean scattering angle is smaller than about 20°. For still larger values of the thickness, d 1/σN, the angular distribution becomes of the form W(Θ) ∝ cos²Θ. The mean angle of scattering Θ then attains its maximum value Θmax ≈ 33°, and remains constant when the thickness increases still further (‘normal’ case, or ‘complete diffusion’). The thickness for which the normal case is reached is called the ‘normal thickness’ dn. Finally, electrons emerge from the foil also on the side of the incident beam. These electrons are either primary electrons which are deflected in the backward direction by single, plural or multiple scattering (back-scattering or back-diffusion), or secondary electrons which, however, are of no interest here. The number of back-scattered electrons reaches a saturation value for a definite thickness dr, the ‘thickness for saturation back-scattering’, or ‘back-diffusion’ thickness.

    1.1 SINGLE SCATTERING

    The probability that an electron with kinetic energy E is scattered during the passage of a foil of thickness d and atomic number Z through an angle Θ into the solid angle dΩ is given by

    where N is the number of scattering atoms per cm³. For the pure Coulomb field of a point charge without shielding we get, according to Mott⁴:

    Here qRuth. is the Rutherford cross-section

    The factor R(E, Z, Θ) calculated by Mott cannot be expressed in analytic form. McKinley and Feshbach⁵ expanded this factor in powers of Zα (α = fine structure constant), and the coefficients have been tabulated up to (Zα)⁴. R is different for electrons and positrons, and this difference increases with increasing energy and increasing angle of scattering and disappears in the non-relativistic limit (R = 1). R is shown in Fig. 1 for 2 MeV electrons and various scattering angles. The expression given above is a good representation of the experimental results for a large range of energy and scattering angle. However, the following deviations should be noted:

    Fig. 1 The ratio R = qMott/qRuth. as a function of αZ for various scattering angles and 2 MeV electrons. Experimental points from Paul and Reich⁶; the value for Al for 60° is fitted to the theoretical value

    (1) For very small scattering angles the shielding of the Coulomb field of the nucleus by the atomic electrons is no longer negligible. The resulting deviations decrease with increasing energy. They have been calculated by Molière⁷ on the basis of the Thomas–Fermi model, and are shown in Fig. 2, where Θ0 is the ‘shielding angle’

    Fig. 2 Influence of the shielding of the Coulomb field by the atomic electrons on the scattering cross-section for various

    Enjoying the preview?
    Page 1 of 1