Abbas, Syed Afsar - Group Theory in Particle, Nuclear, and Hadron physics-CRC (2016)

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 540

GROUP

THEORY IN
PARTICLE, NUCLEAR,
AND HADRON PHYSICS
GROUP
THEORY IN
PARTICLE, NUCLEAR,
AND HADRON PHYSICS

SYED AFSAR ABBAS

c& CRC Press


Taylor & Francis Group
Boca Raton London New York

CRC Press is an imprint of the


Taylor & Francis Group, an informa business
A C H A P M A N & HALL BOOK
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742

© 2017 by Taylor & Francis Group, LLC


CRC Press is an imprint of Taylor & Francis Group, an Informa business

No claim to original U.S. Government works

Printed on acid-free paper


Version Date: 20160222

International Standard Book Number-13: 978-1-4987-0466-3 (Hardback)

This book contains information obtained from authentic and highly regarded sources. Reasonable efforts
have been made to publish reliable data and information, but the author and publisher cannot assume
responsibility for the validity of all materials or the consequences of their use. The authors and publishers
have attempted to trace the copyright holders of all material reproduced in this publication and apologize to
copyright holders if permission to publish in this form has not been obtained. If any copyright material has
not been acknowledged please write and let us know so we may rectify in any future reprint.

Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmit-
ted, or utilized in any form by any electronic, mechanical, or other means, now known or hereafter invented,
including photocopying, microfilming, and recording, or in any information storage or retrieval system,
without written permission from the publishers.

For permission to photocopy or use material electronically from this work, please access www.copyright.
com (http://www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood
Drive, Danvers, MA 01923, 978-750-8400. CCC is a not-for-profit organization that provides licenses and
registration for a variety of users. For organizations that have been granted a photocopy license by the CCC,
a separate system of payment has been arranged.

Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used
only for identification and explanation without intent to infringe.

Library of Congress Cataloging‑in‑Publication Data

Names: Abbas, Syed Afsar, 1946- author.


Title: Group theory in particle, nuclear, and hadron physics / Syed Afsar
Abbas.
Description: Boca Raton, FL : CRC Press, Taylor & Francis Group, [2016] |
©2016 | Includes bibliographical references and index.
Identifiers: LCCN 2016006097| ISBN 9781498704663 (hardback ; alk. paper) |
ISBN 1498704662 (hardback ; alk. paper)
Subjects: LCSH: Group theory. | Particles (Nuclear physics) | Nuclear
physics. | Hadrons.
Classification: LCC QC20.7.G76 A23 2016 | DDC 539.701/5122--dc23
LC record available at http://lccn.loc.gov/2016006097

Visit the Taylor & Francis Web site at


http://www.taylorandfrancis.com

and the CRC Press Web site at


http://www.crcpress.com
To
Ratna
Contents

List of Figures xi

List of Tables xiii

Preface xv

Author xvii

1 Basic Symmetry Concepts 1

1.1 Symmetries Everywhere . . . . . . . . . . . . . . . . . . . . . 1


1.2 Elementary Concepts of Symmetry . . . . . . . . . . . . . . 3
1.3 Kepler’s Construction: A Mistake? . . . . . . . . . . . . . . . 6
1.4 Symmetries and Conservation Laws . . . . . . . . . . . . . . 9
1.5 Appendix A: Mathematics as the Language of Nature . . . . 11
1.6 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . 18

2 Group Theory 19

2.1 Definitions and Examples . . . . . . . . . . . . . . . . . . . . 19


2.2 Subgroups and Cyclic Groups . . . . . . . . . . . . . . . . . 28
2.3 Cosets and Normal Subgroups . . . . . . . . . . . . . . . . . 32
2.4 Factor Group . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.5 Homomorphism and Isomorphism . . . . . . . . . . . . . . . 37
2.6 Torsion Group and Betti Number . . . . . . . . . . . . . . . 40
2.7 Appendix B: Matrices . . . . . . . . . . . . . . . . . . . . . . 48
2.8 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . 54

3 Unitary Symmetry 67

3.1 Lie Groups and Lie Algebras . . . . . . . . . . . . . . . . . . 67


3.2 Some Specific Lie Groups . . . . . . . . . . . . . . . . . . . . 73
3.2.1 SO(2): Special Orthogonal Group in Two Dimensions 73
3.2.2 SO(3): Special Orthogonal Group in Three Dimensions 76
3.2.3 SU(2): Special Unitary Group in Two Dimensions . . 78
3.2.4 SU(3): The Special Unitary Group in Three Dimesions 80
3.3 Adjoint Representation . . . . . . . . . . . . . . . . . . . . . 84

vii
viii Contents

3.4 Adjoint Representations of SU(2) . . . . . . . . . . . . . . . 86


3.5 Cartan Subalgebra and Roots/Weight Space . . . . . . . . . 87
3.6 Fractional Charges in SU(n) . . . . . . . . . . . . . . . . . . 93
3.7 Appendix C: Linear Vector Space . . . . . . . . . . . . . . . 100
3.8 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . 108

4 Permutation Symmetry 117

4.1 Indistinguishable Particles in Quantum Mechanics . . . . . . 117


4.2 Transpositions and Permutations . . . . . . . . . . . . . . . . 118
4.3 Disjoint Cycles and Signatures . . . . . . . . . . . . . . . . . 122
4.4 Cayley’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . 126
4.5 Symmetric Group S3 . . . . . . . . . . . . . . . . . . . . . . 126
4.6 Classes and Partitions . . . . . . . . . . . . . . . . . . . . . . 130
4.7 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . 132

5 Young Diagram 139

5.1 Symmetries and Young Diagrams . . . . . . . . . . . . . . . 139


5.2 Young Operators . . . . . . . . . . . . . . . . . . . . . . . . . 146
5.3 Young Diagram and Unitary Symmetry . . . . . . . . . . . . 151
5.4 Product of Irreducible Representations . . . . . . . . . . . . 160
5.5 Multiplets of the SU (n − 1) Subgroup of SU (n) . . . . . . . 162
5.6 The Reduction SU (m + n) → SU (m) ⊗ SU (n) . . . . . . . . 165
5.7 Reduction of SU (mn) → SU (m) ⊗ SU (n) . . . . . . . . . . . 168
5.8 Coefficients of Fractional Parentage and Isoscalar Factors . . 171
5.9 Appendix D: Representation Theory . . . . . . . . . . . . . . 176
5.10 Appendix E: Symmetry in Quantum Mechanics . . . . . . . 184
5.11 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . 189

6 Quantum Chromodynamics 203

6.1 Why the SU (3)c Group? . . . . . . . . . . . . . . . . . . . . 203


6.2 QCD Langrangian and Asymptotic Freedom . . . . . . . . . 206
6.3 Colour Singlet States and Confinement . . . . . . . . . . . . 211
6.4 Pentaquarks . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
6.5 Large Nc Colour QCD . . . . . . . . . . . . . . . . . . . . . 220
6.6 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . 224

7 Quark Model 227

7.1 Current and Constituent Quarks . . . . . . . . . . . . . . . . 227


7.2 The Eightfold Way Model . . . . . . . . . . . . . . . . . . . . 234
7.3 SU (3)F Flavour Model . . . . . . . . . . . . . . . . . . . . . 239
7.4 Quark Model Calculations . . . . . . . . . . . . . . . . . . . 249
7.5 SU (6)SF Model . . . . . . . . . . . . . . . . . . . . . . . . . 259
Contents ix

7.6 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . 264

8 Bag Models 271

8.1 Why a Bag? . . . . . . . . . . . . . . . . . . . . . . . . . . . 271


8.2 Confinement in a Spherically Static Bag . . . . . . . . . . . . 272
8.3 MIT Bag Model . . . . . . . . . . . . . . . . . . . . . . . . . 275
8.4 Finite Mass Quarks in a Bag . . . . . . . . . . . . . . . . . . 292
8.4.1 Magnetic Moments . . . . . . . . . . . . . . . . . . . . 295
8.4.2 Axial Vector Coupling Constant . . . . . . . . . . . . 296
8.4.3 Spin Structure of the Nucleon . . . . . . . . . . . . . . 297
8.5 Scalar and Vector Confining Potentials . . . . . . . . . . . . 300
8.6 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . 303

9 Hadron Physics 309

9.1 Harmonic Oscillator Model . . . . . . . . . . . . . . . . . . . 309


9.2 Configuration Mixing . . . . . . . . . . . . . . . . . . . . . . 317
9.3 Deformed Nucleon . . . . . . . . . . . . . . . . . . . . . . . . 323
9.4 Weak Charges . . . . . . . . . . . . . . . . . . . . . . . . . . 325
9.5 Magnetic Moments . . . . . . . . . . . . . . . . . . . . . . . 333
9.6 Spin of Nucleon in the Quark Model . . . . . . . . . . . . . . 339
9.6.1 Spin of a Deformed Nucleon . . . . . . . . . . . . . . . 340
9.6.2 Spin of Nucleon with Configuration Mixed Wave Func-
tion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348
9.7 Colour Confinement in QCD and Deformed Baryons . . . . . 356
9.8 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . 359

10 Glashow-Salam-Weinberg Model 367

10.1 Its Chiral and Non-Chiral Structure . . . . . . . . . . . . . . 367


10.2 Spontaneous Symmetry Breaking . . . . . . . . . . . . . . . 372
10.3 GSW Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 374
10.4 Millicharged Neutrinos . . . . . . . . . . . . . . . . . . . . . 380
10.5 Second Generation Fermions . . . . . . . . . . . . . . . . . . 381
10.6 Right-Handed Neutrino . . . . . . . . . . . . . . . . . . . . . 383
10.7 Gauge Bosons and Neutral Currents . . . . . . . . . . . . . . 386
10.8 Neutrino Oscillation . . . . . . . . . . . . . . . . . . . . . . . 389
10.9 Appendix F: Lorentz and Poincare Groups . . . . . . . . . . 399
10.10Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . 407

11 Symmetry in Nuclei 411

11.1 Isospsin Symmetry in Nuclei . . . . . . . . . . . . . . . . . . 411


11.2 SU (4) Symmetry and Saturation in Nuclei . . . . . . . . . . 418
11.3 Distinguishable Protons and Neutrons in Nuclei . . . . . . . 420
x Contents

11.4 Gamow-Teller Strengths in Nuclei . . . . . . . . . . . . . . . 423


11.5 SU (2)A Nusospin Symmetry in Nuclei . . . . . . . . . . . . . 426
11.6 Quantum Groups in Nuclei . . . . . . . . . . . . . . . . . . . 431

12 Quarks in Nuclei 437

12.1 The EMC Effect and the Nucleus . . . . . . . . . . . . . . . 437


12.2 Hidden Colour in Multiquarks in Nuclei . . . . . . . . . . . . 438
12.3 Quarks in A=3 Nuclei . . . . . . . . . . . . . . . . . . . . . . 448
12.4 ∆ Excitations in the Nucleus . . . . . . . . . . . . . . . . . . 449
12.5 M1 Strength in Nuclei . . . . . . . . . . . . . . . . . . . . . . 451
12.6 Gamow-Teller (GT) Strength in Nuclei . . . . . . . . . . . . 457
12.7 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . 458

13 Quark Gluon Plasma(QGP) 463

13.1 Basics for QGP . . . . . . . . . . . . . . . . . . . . . . . . . 463


13.2 Finite Lie Group Transformations . . . . . . . . . . . . . . . 467
13.3 Group Characters of the Lie Group . . . . . . . . . . . . . . 471
13.4 Measure Function of SU(n) . . . . . . . . . . . . . . . . . . . 476
13.5 Symmetries and Partition Functions . . . . . . . . . . . . . . 478
13.6 Colour Singlet and Coloured QGP States . . . . . . . . . . . 480
13.7 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . 488

14 Topology for Hadrons 493

14.1 Why Topology? . . . . . . . . . . . . . . . . . . . . . . . . . 493


14.2 Homotopy Groups . . . . . . . . . . . . . . . . . . . . . . . . 495
14.3 Linear and Non-Linear Sigma Models . . . . . . . . . . . . . 498
14.4 Skyrme Model . . . . . . . . . . . . . . . . . . . . . . . . . . 500
14.5 SU(3) Adjoint Representation Skyrme Model . . . . . . . . . 503
14.6 Appendix G: Introduction to Topology . . . . . . . . . . . . 505
14.7 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . 513

Bibliography 515

Index 523
List of Figures

1.1 Reflection symmetry of a neck-tie and symmetry of an equi-


lateral triangle under rotation . . . . . . . . . . . . . . . . . 4
1.2 Rotation and reflection symmetry of a dumbbell and an irreg-
ular figure . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Symmetry of cyclic groups C4 (a), C3 (b), C2 (c), C1 (d) . . . 5
1.4 Rotation and reflection symmetry of the dihedral group DN 6
1.5 The five Platonic solids . . . . . . . . . . . . . . . . . . . . . 7
1.6 Two identical particles with a two–body potential . . . . . . 10

2.1 Symmetries of an equilateral triangle and a square . . . . . 23


2.2 Lattice diagram of the subgroups of the groups Z4 , V and D3 30
2.3 Lattice diagram showing the subgroup structure of D4 . . . 57

3.1 Rotation of a two-dimensional vector – SO(2) group . . . . 73


3.2 Root-vector representation of the adjoint representation of
SU (2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
3.3 Weight vectors of the j-representation of SU (2) . . . . . . . 91
3.4 Weights of the baryon octet states in SU (3) . . . . . . . . . 92
3.5 Roots of the baryon octet states in SU (3) . . . . . . . . . . 94
3.6 Passive transformation on a two-dimensional vector . . . . . 102

4.1 Regular tetrahedron symmetry as isomorphic to the symmet-


ric group S4 . . . . . . . . . . . . . . . . . . . . . . . . . . . 129

6.1 Schematic plot of the ratio R which proves the presence of


three colours . . . . . . . . . . . . . . . . . . . . . . . . . . . 204

7.1 Infinite momentum frame in deep inelastic scattering . . . . 229


7.2 Spin 1/2 baryons plus pseudoscalar and vector meson octets 243
7.3 Spin 3/2 baryon decuptet . . . . . . . . . . . . . . . . . . . 244

9.1 Jacobi coordinates for three-dimensional space . . . . . . . . 312


9.2 Baryon spectrum under the lowest-order perturbation theory 320
9.3 Locations of the three quarks at the positions 1, 2 and 3 . . 357

xi
xii List of Figures

11.1 One and two triton separation energies as a function of triton


number . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 430

12.1 Experimentally determined central hole in the charge density


distribution of A=3 and A=4 nuclei and the standard theo-
retical expectation . . . . . . . . . . . . . . . . . . . . . . . 441

13.1 Schematic picture of high energy heavy ion collisons to create


QGP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 465
13.2 Weight diagram of the triplet, antitriplet, and octet represen-
tations used to calculate partition functions . . . . . . . . . 481
13.3 Def f for singlet, octet, and 27-plet representation for two
flavours . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 485

14.1 Topological mapping and winding numbers . . . . . . . . . . 497


14.2 Open interval on a real number line . . . . . . . . . . . . . . 506
14.3 An open set on a real number line . . . . . . . . . . . . . . . 506
14.4 An open ball . . . . . . . . . . . . . . . . . . . . . . . . . . . 507
14.5 Path-connected points . . . . . . . . . . . . . . . . . . . . . 508
14.6 Hausdorff space . . . . . . . . . . . . . . . . . . . . . . . . . 508
14.7 Square and circle topology . . . . . . . . . . . . . . . . . . . 509
14.8 Euler number for a cube and a tetrahedron . . . . . . . . . 510
14.9 Tetrahedron on a sphere . . . . . . . . . . . . . . . . . . . . 511
14.10 Betti number of networks . . . . . . . . . . . . . . . . . . . 511
14.11 Euler number of S 2 and a torus . . . . . . . . . . . . . . . . 514
List of Tables

1.1 Invariance and conservation laws connection . . . . . . . . . 11

2.1 Composition table of the cube roots of unity . . . . . . . . . 22


2.2 Composition table of the cyclic group C3 . . . . . . . . . . . 24
2.3 Composition table for the group D3 . . . . . . . . . . . . . . 24
2.4 Reflections and rotations as blocks in the D3 group . . . . . 25
2.5 Composition table of D4 . . . . . . . . . . . . . . . . . . . . 26
2.6 Composition table of addition modulo 4 . . . . . . . . . . . 27
2.7 Viergruppe V or 4-group . . . . . . . . . . . . . . . . . . . . 28
2.8 Factor group of D3 . . . . . . . . . . . . . . . . . . . . . . . 35
2.9 C2 and Z2 composition table . . . . . . . . . . . . . . . . . . 35
2.10 Z3 composition table . . . . . . . . . . . . . . . . . . . . . . 37
2.11 Order three group . . . . . . . . . . . . . . . . . . . . . . . . 38
2.12 Composition table of square roots of unity . . . . . . . . . . 55
2.13 Composition table of fourth roots of unity . . . . . . . . . . 55
2.14 Z6 the addition modulo 6 composition table . . . . . . . . . 56
2.15 Left cosets of {0,3} of Z6 . . . . . . . . . . . . . . . . . . . . 58
2.16 Viergruppe V rearranged . . . . . . . . . . . . . . . . . . . . 59
2.17 Factor group of V . . . . . . . . . . . . . . . . . . . . . . . . 59
2.18 Factor group of D4 . . . . . . . . . . . . . . . . . . . . . . . 60
2.19 Blocked composition table of D4 . . . . . . . . . . . . . . . 60

3.1 Fractional charges of h-electrons in the groups SU (2)h2 to


SU (4)h4 as in FQHE . . . . . . . . . . . . . . . . . . . . . . 100

4.1 Composition table for the symmetric group S3 . . . . . . . . 127


4.2 Number of elements in classes of S4 and S5 . . . . . . . . . 137

6.1 Hidden colour components in the multiquark systems . . . . 217


+
7.1 Parallel structure of 12 baryons and 0− mesons leading to the
eightfold way model . . . . . . . . . . . . . . . . . . . . . . . 235
7.2 The complete set of SU(3) commutation relations . . . . . . 236
7.3 Mixed symmetric wave functions for the spin-1/2 octet
baryons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247

xiii
xiv List of Tables

7.4 Magnetic moments of spin-1/2 baryon octet in the quark


model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256

8.1 Eigenvalues in a central potential . . . . . . . . . . . . . . . 273


8.2 Experiment [E142] compared to predictions of MIT bag model
(for massless quarks w0 = 2.04 and for m 6= 0, ER=2.61, and
mR=1.01); the mR → ∞ case corresponds to non-relativistic
quark model; best configuration mixed cases (1) and (2); and
the deformed nucleon model case . . . . . . . . . . . . . . . 299

9.1 Baryon space wave functions in the Harmonic Oscillator


Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 314
9.2 Symmetric SU (6)F S wave functions for N and ∆, octet and
decuptet only for spins, J = 12 and 32 , respectively . . . . . . 317
9.3 Semi-leptonic decays of baryons with deformation (only single
parameter for all these fits) . . . . . . . . . . . . . . . . . . 332
9.4 Results for u↑↓ and d↑↓ terms arising from different parts of
the configuration mixed quark model wave functions . . . . 354
9.5 Polarized structure function values of different quantities –
the experimental values (E); the configuration mixed quark
model values (1 and 2, Equation 9.151); and the deformed
nucleon results (D) . . . . . . . . . . . . . . . . . . . . . . . 355

11.1 Inter-triton cluster bond energies of neutron-rich nuclei . . . 427


11.2 Experimental and theoretical data for Superdeformed Bands
in 194 Hg(2) compared with our results . . . . . . . . . . . . 435
11.3 Supersymmetry in nucleus – fitting of identical Superde-
formed Bands in the neighbouring nuclei, 152 Dy and 151 T b∗ 436

12.1 Colour singlet components in 6-, 9- and 12-quark systems -


the rest are hidden colour . . . . . . . . . . . . . . . . . . . 448
Preface

This book is primarily designed for those who are embarking on a research
career in particle, nuclear or hadron physics. The single-minded goal is to
present group theory in the manner it is practically applied in these disci-
plines in contemporary research. Group theory is presented here in a simple,
consistent and unified fashion, so that the field does not appear too threaten-
ing to the novice. However, sufficient advanced material has also been supplied
in this work, so as to provide a source of valuable information and present tan-
talizing challenges to the experienced researcher as well. Though the book is
targeted at theorists, the direct and lucid approach adopted throughout may
very well render it accessible to experimentalists too.
A unique feature of this book is the large number of solved problems, in
addition to the few unsolved ones. The detailed individual mathematical steps
and corresponding physical implications are provided in the main text, as well
as in the pertinent “Problem Solutions” sections. Hence, this work may be
utilized within a formal lecture framework (including those for B.Sc., B.Tech.,
M.Sc., M.Tech. and M.Phil. degrees), as well as for individual self-study.
I am thankful to Ms. Aastha Sharma of CRC Press for her continuous encour-
agement while the book was under preparation. Timely assistance from Mr.
S. M. Hasnain is gratefully acknowledged. Thanks are due to Vipin and Ra-
mana Babu for help in typing. Critical reading, careful editing and aesthetic
embellishment of the manuscript by Dr. Samar Abbas is worthy of recog-
nition. Encouragement from Dr. Ummi Abbas and Dr. Alex Sozzetti is also
appreciated. I am especially indebted to my wife for her limitless patience and
constant support throughout this Tughlaqian endeavour.

Syed Afsar Abbas


Jafar Sadiq Research Institute, Aligarh, India

xv
Author

- M.Sc. (Physics), McMaster University, Canada

- Ph.D. (Physics), Rutgers University, USA

- Sir Charles Clore (Distinguished) Postdoctoral Fellow, Weizmann Institute


of Science, Israel

- Assistant Professor (BMFT), Institut fuer Kernphysik, Technische Univer-


sitaet, Darmstadt, Germany

- Senior Fellow, Dept. of Theoretical Physics, University of Manchester, UK

- Professor, Institute of Physics, Bhubaneswar, India

- UGC Visiting Professor, Centre for Theoretical Physics, New Delhi, India

- UGC Visiting Professor, Physics Dept., Aligarh Muslim University, India

- (Founding) Director, Centre for Interdisciplinary Research in Basic Sciences,


New Delhi, India

xvii
Chapter 1
Basic Symmetry Concepts

1.1 Symmetries Everywhere . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1


1.2 Elementary Concepts of Symmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Kepler’s Construction: A Mistake? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.4 Symmetries and Conservation Laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.5 Appendix A: Mathematics as the Language of Nature . . . . . . . . . . 11
1.6 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

1.1 Symmetries Everywhere


Prior to embarking on our odyssey to comprehend the abstract expressions
and finer intricacies of group theory, we shall first analyze the simpler concepts
of symmetry.
Now, humans are endowed with an intuitive understanding of what “sym-
metry”denotes. All around we are literally surrounded by objects – some dis-
play no features of symmetry whatsoever, while others exhibit clear evidence
of symmetry that is intrinsically recognized by humans [1]. In fact, during the
course of our evolution as a species, symmetry has grown intimately inter-
twined with the very concept of beauty itself [2]. Whatever is symmetric in
nature appears beautiful to the human eye.
Little wonder then, that symmetry manifests itself in all fields of human
creative enterprise, be it in art, architecture, music or poetry. A few prominent
examples from architecture would suffice to illustrate this point:

- From about 2500 B.C. to 500 B.C. the Cyclopean Ziggurats of Ur and the
Gargantuan Towers of Babylon in Ancient Iraq which transfixed all travellers
of yore were gloriously symmetrical structures testifying to the architectural
genius of Greater Middle Eastern civilization.

- The Blue Mosque of Istanbul in Turkey (1609 – 1617), also known as


the Sultan Ahmet Mosque, was built by the master Turco-Islamic architect
Mehmet during the magnificent Ottoman Caliphate. A splendid structure tow-
ering over the mighty Bosphorus, it is the only Istanbul mosque containing
six symmetrically placed minarets – in marked contrast to the more common

1
2 Group Theory in Particle, Nuclear, and Hadron Physics

quartet.

- The Peruvudaiyar Kovil Temple in Thanjavur, Tamil Nadu (1010 AD)


is the tallest temple in the world. A masterpiece of the highly distinctive and
monumentalist Dravida school of architecture, it exhibits countless striking
symmetries, right from the large-scale architecture of its colossal Gopuram to
the intricate carvings decorating its wonderful walls. This masterpiece demon-
strates the hoary antiquity of the indigenous Shaivite religion, as well as the
astonishing continuity of the present Dravidian civilization.

- The Russian Cathedral of St. Basil the Blessed in the Red Square of
Moscow is a most remarkable and eye-catching structure of unique beauty. The
bright enlivening colours which appear interwoven into an intricate tapestry
of geometrical symmetry has struck visitors as one of the most breathtaking
and awe-inspiring creations of the Oriental Slavo-Byzantine civilization [2].

- Finally, mention must be made of the glorious Taj Mahal of Agra in


Oudh [3], constructed by the powerful rulers of the Later Timurid Caliphate,
who are occasionally mistakenly termed as “Mughals”. Often considered the
most beautiful structure ever created by man, various layers of symmetry are
evident right from the opulent Islamic domes towering into the Awadhi sky,
through the whirling Perso-Arabic inscriptions adorning the translucent mar-
ble walls, down to the precious Kashmiri sapphires, Rajastani emeralds and
Golcunda diamonds which once adorned its splendid surface.

In the realm of decorative arts, one may mention the works of the prolific
Nordic-Dutch lithographer Maurice Escher [4]. Including dazzling outpourings
of geometrical splendour exhibiting a perplexing interplay of different symme-
tries and optical illusions such as “Air and Water” (1938) or “Ascending and
Descending” (1960), his Netherlandish productions have a special appeal for
physicists and mathematicians.
The performing arts also exhibit a marked fondness for symmetry. For
example, poets have always waxed eloquent about the beauty of symmetry
and have always turned lyrical regarding the rhythm of rhyme. For example,
a poem written by Su Dongpo (eleventh century) in China, consisting of eight
vertical lines of seven characters each. The poem can be read forward and
backward each time with correct rhyme and meter. In this context, one must
remember that the hoary Chinese language is read commencing from right,
top-down a column, and then down along the second column and so forth.
Musicians too are not far behind in seeking beauty from symmetry. For
example, the Crab Canon of the Nordic-German musician J. S. Bach (seven-
teenth century) is a violin duet where each voilin’s music is a time reversed-
version of that of the other violin.
So clearly symmetry and its intrinsic beauty has been the focus of much
human creative endeavour among artists. The same can be said of scientists
Basic Symmetry Concepts 3

too. The first motivation, more as an act of intellectual satisfaction, is for


seeking beauty for its own sake
Let us recollect the Nordic-British poet John Keats’ final lines in his 1819
poem “Ode on a Grecian Urn”

Beauty is truth, truth beauty, that is all


Ye know on earth, and all ye need to know.

A further quote from Paul Dirac serves to emphasise the fascination physi-
cists have with symmetry [5]: “It seems that if one is working from the point
of view of getting beauty in one’s equations, and if one has really a sound
insight, one is on a sure line of progress.”
In fact, the search for symmetry in nature – with symmetry being sought
for its own sake – has been found to be an extremely fruitful enterprise by
scientists. First we study symmetry as a scientific concept, and thereafter move
over to group theory, which is the mathematical language that incorporates
symmetry in a consistent manner.

1.2 Elementary Concepts of Symmetry


To understand symmetry, let us consider figures drawn on a page. Sym-
metry of a figure exists if the shape it represents does not change in size or
layout when subjected to a rotation or reflection within itself. These changes
are termed transformations. A “no-change” is indeed a transformation of a
special kind itself, and is denominated as an identity. Just as multiplication
by unity does not alter the value of a quantity, so in the case of symmetry,
the value of symmetry does not change when the identity participates.
In Figure 1.1(a),(b), the figure (denoting a neck – tie) before and after the
transformation appears the same. Hence, it is symmetric under the reflection
transformation.

Note that a rotation by = 2π returns Figure 1.1(c) to the original con-
3
figuration. So the cycle is repeated. All the more reason that the “identity” is
a transformation. Note that points 1, 2, 3 are external labels (not existing on
the figure itself) and hence the figures under all the above three transforma-
tions Figure 1.1(c),(d),(e) – appear the same, or are in fact identical. Hence
these entities are symmetric under these transformations.
Figure 1.2(a) provides a rough idea of symmetry. Essentially, symmetry
involves some entity. This entity retains its identity when a transformation
(namely, a series of well-defined changes) is imposed upon it. It is hence stated
that the object is symmetric under that particular transformation.
Of course there could be a large number of figures that may appear to have
4 Group Theory in Particle, Nuclear, and Hadron Physics

(a] (b)

(c) (d)

\2 1, 1 3,

3 2
(e)

3 2,

FIGURE 1.1: Reflection symmetry of a neck-tie and symmetry of an equi-


lateral triangle under rotation

m
l

(a)

m2

(b)

FIGURE 1.2: Rotation and reflection symmetry of a dumbbell and an ir-


regular figure
Basic Symmetry Concepts 5

2 1

(a)

3 4

2 1

(b)

(c)

(d)

FIGURE 1.3: Symmetry of cyclic groups C4 (a), C3 (b), C2 (c), C1 (d)

no symmetry at all. Does it then imply that irregular figures do not have any
symmetry?
You may say no. But we have already defined “identity” as an operation
that is a valid transformation. So even the most asymmetrical figure has a
symmetry - it is symmetric to itself, and the example is Figure 1.2(b).
Let us define a few simple symmetry types which will have relevance as
point groups in crystallography and molecular physics.
Symmetry structure CN are figures that have an exactly N -rotation sym-
metry and no mirror symmetry. It is thus called a cyclic group. Cyclic, as it
repeats itself after a full cycle. A few simple examples are given in Figure 1.3.
Now we have given a name to irregular shapes with no apparent symmetry.
They too have a symmetry C1 , i.e., symmetric to themselves. So you rotate
it back through any point, e.g., Figure 1.3(d).
Next the larger symmetry structure DN , the “Dihedral Group”, has ex-
actly N-rotation symmetries (which CN already had) and additional mirror
reflection symmetries.
D1 is called the bilateral symmetry. Our body displays it.
More on D4 (Figure 1.4(e)). Here a total of 8 symmetries (you may figure
these out).

————————————————–
Problem 1.1: Determine D3 for a regular triangle and D5 for a regular
6 Group Theory in Particle, Nuclear, and Hadron Physics
(a) (b) (c)
D4 D
3 D,
2 1
1
2

3 4 3

d) D
i

m2
1
2

(e)
m
3

4
3

m
i m4

FIGURE 1.4: Rotation and reflection symmetry of the dihedral group DN

pentagon. Generalize for a regular n-gonal figure. (Regular n-gonal figure:


closed figure of n sides that are all equal.)
————————————————–
Problem 1.2: We saw that for N -gon figures the identity belongs to the
CN subgroup. Why is the identity not a no-reflection?
————————————————–

1.3 Kepler’s Construction: A Mistake?


The reader, having gained an appreciation of symmetry, should note that
the concept of symmetry has been a motivating factor for the construction
of theories in physical science since time immemorial. The Ionian philosopher
Aristotle (150 BC) attempted to understand the position of planets arising
from the theory of numbers in the framework of the “Harmony of Spheres”.
Moving over from the above two-dimensional symmetry to three-
dimensional symmetry, it turns out that there are exactly five regular solids
exhibiting maximal symmetry. These are termed Platonic solids (named after
the Ionian philosopher Plato) and are depicted in Figure 1.5.
Basic Symmetry Concepts 7

Regular Cube
Tetrahedron (Pyramid)

Octahedron

Dodecahedron Icosahedron

FIGURE 1.5: The five Platonic solids


8 Group Theory in Particle, Nuclear, and Hadron Physics

Hence, the Greeks thought that as these five Platonic solids appear to be
the most symmetric entities in three dimensions, then these should be basic
too. And as such nature may reflect these in its intrinsic structure as well.
Hence several ancient scientists attempted to fit the motion of the six planets
visible to the naked eye (Mercury, Venus, Earth, Mars, Jupiter, and Saturn)
as a projection of the fundamental nature of the Platonic solids.
This was a completely scientific attitude. If there are symmetries known to
us, then fitting the laws of nature within the framework of these is indeed good
science. Today we do the same thing. We try to seek symmetry in physical
laws and physical effects. Or, knowing a particular symmetry, we attempt to
use it to explain the known physical situation and make predictions for some
putative unknown “reality”. Of course, experiments, as the ultimate arbiter
of what is correct and what is wrong, provide the final judgement for such
theories.
However, it is typically prudent to wait for multiple verification and fi-
nal confirmation of experimental results, for there exist numerous instances
where theoreticians have been served veritable Barmecide feasts by perhaps
overenthusiastic experimentalists. Examples of potentially pathbreaking re-
sults which had to be withdrawn under closer scrutiny include the curious
case of the psuedo-detection of the theta particle as pentaquark, the pur-
ported sighting of planet Vulcan, the reported achievement of cold fusion,
and the supposed discovery of gravitational waves. Some of these empirical
mirages are discussed later on in this work.
The sixteenth century Nordic-Dutch astronomer Tycho Brahe had com-
pleted a meticulous observational record of planetary motion without the aid
of a telescope. The data was passed on to his ingenious assistant Johannes
Kepler. Kepler used this data to arrive at his famous laws of planetary mo-
tion – the so-called Kepler’s laws of planetary motion. We know that it was
these laws which described the reality of planetary motion quite well. Subse-
quently, these were used by the Nordic-British scientist Newton to arrive at
his celebrated theoretical Law of Gravitation.
However, note that Kepler did not arrive at these laws of planetary motion
in a smooth manner. He had inherited the above-mentioned Greek concept
(should it be termed a prejudice?) of symmetry wherein Platonic solids would
form a fundamental basis for the description of natural reality, whence the
“Heavens” should be the ideal ground for such an interplay. He used Tycho
Brahe’s data to devise laws of motion of planets within the framework of
Platonic solids. Thus he endeavoured to utilize this data to fit the motion
of the six known planets, with the five Platonic solids “inscribed” between
them. Hence, starting with the Sun at the centre, he fitted the motion of
Mercury with one Platonic solid tangential to Mercury’s orbit and on whose
vertices there would be another spherical surface, on which the movements
of Venus would be constrained. The last one was Saturn, which moved in a
circular orbit on a sphere tangential to a “regular cube” Platonic solid. On
the circle inscribed on it moved Jupiter and so forth. This novel model was
Basic Symmetry Concepts 9

actually published by Kepler as his construction of the planetary motion as


“Mysterium Cosmographicum” in 1595.
We know today that this was completely wrong. Fortunately Kepler, being
a true scientist, was able to correct for errors in his own logic and was thus
able to realize the significance of the elliptical orbits in planetary motion.
With this new insight, he rejected his earlier model, and ultimately arrived at
his “correct” laws of planetary motion which are known to us today.
The moral of the above story is that first, the fact that one is using some
symmetry principle or some fancy mathematics is no guarantee for the correct-
ness of the theory. For the theory to be justified one has to use the “correct”
symmetry. Mathematics is a language and as such, one has to exploit the cor-
rect language to describe a particular physical phenomenon. One cannot use
the word “right hand” for something that is actually “left hand”. But how do
we know which is the correct language? Well, our experimental colleagues will
provide the final judgement by confirming or rejecting a theory by performing
appropriate experiments.

1.4 Symmetries and Conservation Laws


As we have seen above, it is an amazing fact of nature that physical systems
possess symmetries and which arise due to it being invariant under some
suitable transformations. So for example we found that an ideal equilateral
triangle has a cyclic C3 symmetry due to it being invariant (did not change)
under suitable rotations around its naturally defined symmetry axis. Another
amazing fact of nature is that there exist conservation laws. Among the many
properties of physical systems that change with time, a few remain constant
thereby leading to conservation laws. We find that whatever the process be,
the total electric charge before a particular reaction and after it remains the
same. Therefore, we say that the electric charge is conserved. We know of
other conservation laws of energy, linear momentum, angular momentum, etc.
These conservation laws allow us to understand which particles and which
processes may occur and under what specific physical circumstances.
A most interesting aspect of physical reality is that the symmetries and
the conservation laws are intrinsically connected.
For example, we know that space is homogeneous. It means that we can
be anywhere in space, be in Aligarh, Delhi, Kolkata, Turin, physical laws
will remain the same. So to say, there is no origin in space. Therefore the
absolute position in space is unmeasurable. This means that the system is
symmetric under translation. Now that we have identified the symmetry (that
of translation) what is the conservation law associated with it?
Consider an isolated system of two identical particles 1 and 2 in Figure 1.6.
With respect to some origin, the potential energy is specified as V (r~1 , r~2 ). Let
10 Group Theory in Particle, Nuclear, and Hadron Physics

1
2

rT
2
r
1

R2

r1 0

0'

FIGURE 1.6: Two identical particles with a two–body potential

us translate the origin by −a. ~ With respect to this new origin, the potential
energy is given as V (r~1 0 , r~2 0 ) with r~1 0 = r~1 + ~a and r~2 0 = r~2 + ~a.
Now because of translational invariance it does not matter where the origin
is. So,

V (r~1 0 , r~2 0 ) = V (r~1 + ~a, r~2 + ~a) (1.1)


this means that V should be a function of (r~1 − r~2 ). Note that,r~1 0 − r~2 0 =
(r~1 + ~a − r~2 + ~a) = r~1 − r~2 .
Hence the above equation for the potential will hold if V is a function of
relative coordinates r~1 − r~2 , that is, V (r~1 − r~2 ).
Now, the force is

∂V (r~1 − r~2 ) ∂V (r~1 − r~2 )


F~ = F~1 + F~2 = − −
∂ r~1 ∂ r~2
∂V (r~1 − r~2 ) ∂V (x) ∂x −∂V (x)
− =− =
∂ r~1 ∂x ∂ r~1 ∂x
∂V (r~1 − r~2 ) ∂V (x) ∂x ∂V (x)
− =− =
∂ r~2 ∂x ∂ r~2 ∂x
So,
∂V (x) ∂V (x)
F~ = − + =0 (1.2)
∂x ∂x
d~
p
F~ = = 0 p~ = constant (1.3)
dt
Therefore, momentum is constant.
Hence unmeasurability of the absolute position implies symmetry or in-
variance under translation, which in turns implies Conservation of Linear Mo-
mentum.
The above connection as: “unmeasurabiity of some physical property →
invariance under some physical transformation → some conseravation law” is
Basic Symmetry Concepts 11
TABLE 1.1: Invariance and conservation laws connection

UNMEASURABILITY SYMMETRY LEADS TO EXACT


OF IMPLIED CONSERVATION OR NOT

Absolute position Translation Linear Yes


(homogeneity) momentum
Absolute direction Rotation Angular Yes
(isotropy) momentum
Absolute origin Time Energy Yes
of time Translation
Difference between Permutation FD,BE Yes
identical particles Symmetry statistics
Indistinguishability of ~r → −~r Parity violated
left or right (Parity) weak int
Relative phase between Baryon gauge Baryon
baryon numbers ψ → eiBα ψ number Yes
Relative phase charged Charge gauge
neutral particles ψ → eiQθ ψ Charge Yes
Difference in coherent ψ → Uψ Isospin group Violated
mix of p and n U in SU (2)I SU (2)I electromag.
Difference in coherent Ψ → UΨ Flavour group Broken
mix of u,d,s quarks U in SU (3)F SU (3)F med-strong
Difference in coherent φ → Uφ QCD group Exact
mix of R,B,G colours U in SU (3)c SU (3)c

well obeyed by our physical reality, both classical and quantum. We give a
partial list in Table 1.1 here. A bigger list can be found in [6] and [7].

—————————————————————-
Problem 1.3: Show that for a system describable by a classical La-
grangian, unmeasurability of absolute position implies that the linear mo-
mentum is conserved.
—————————————————————–

1.5 Appendix A: Mathematics as the Language of Na-


ture
(a) Hoary Tradition
In this section we investigate how the field of applied linguisics can assist
12 Group Theory in Particle, Nuclear, and Hadron Physics

in understanding the importance of mathematics in general, and in compre-


hending its relevance for the study of nature in particular. Remember that
group theory is the core of how mathematics is used in physics today.
The applicability of linguistics to the field of mathematics is underlined
by a prevalent scientific tradition which views mathematics as a – if not the
– language of nature. According to this belief, mathematics is the intrinsic
language of scientists, just as German is the mother tongue of the residents of
Berlin and Arabic is the native speech of the inhabitants of Cairo. This deep-
rooted perception has persisted through the centuries, finding expression in
the hoary writings of the brightest exponents of the field. For example, Galileo
Galilie vividly observed [8].
“Philosophy (i.e., physics) is written in this grand book – I mean the
universe – which stands continually open to our gaze, but it cannot be under-
stood unless one first l earns to comprehend the language and interpret the
characters in which it is written. It is written in the language of mathematics,
and its characters are triangles, circles, and other geometrical figures, without
which it is humanly impossible to understand a single word of it; without
these, one is wandering around in a dark labyrinth”.
In the same vein, Bertrand Russell noted, “Ordinary language is totally
unsuited for expressing what physics really asserts, since the words of everyday
life are not sufficiently abstract. Only mathematics and mathematical logic can
say as little as the physicist means to say” [[9], p.56 – 57].
In fact, a section of the scientific community has taken this view to an
extreme, viewing the whole of nature itself as a creation of mathematics. As
an example, one may recollect the enthusiastic exclamation of James Jeans
that “God is a mathematician” [10].
These instances serve to demonstrate the perseverance of the common
scientific and philosophical belief that mathematics is a language of nature.
(b) Linguistics Applied to Mathematics
Now, if mathematics is indeed a language of nature, or a collection of lan-
guages describing nature, then it logically follows that the field of linguistics
should be emininently useful in understanding mathematics. Viewing mathe-
matics itself as a set of languages and applying principles of linguistics should
then lead to further insights and, perhaps, new discoveries.
To start with, let us see how conventional linguistics defines a language.
According to Noam Chomsky, “A language is a set (finite or infinite) of sen-
tences, each finite in length and constructed out of a finite set of elements”
[11].
Meanwhile, Trager defines a language as follows:
“A language is a system of arbitrary vocal symbols by means of which the
members of a society interact in terms of their total culture” [12].
Infant Linguistics and Mathematics: Within the field of language
studies, the acquisition of linguistic skills by growing infants in order to com-
municate with society at broad provides fascinating parallels to the develop-
Basic Symmetry Concepts 13

ment of mathematical skills by scientists attempting to describe nature at


large.
Human vocal cords being capable of producing a very large range of audible
sounds varying in terms of wavelength and pitch, a growing child at first
experiments with these across the entire spectrum. During the initial phases
of self-absorbed internal development, the child communicates with itself in
sounds that are totally incomprehensible to other humans. However, as the
child then expands its sphere of communication to the outside world, it quickly
learns that only a finite set of sounds is acceptable to the social group it
belongs to. A process of gradual adaptation takes place, with the child learning
to reproduce only the limited group of sounds and words that society imposes.
Now, this process of the linguistic comprehension of proper intelligible
words can be seen as the counterpart of the process of scientific discovery of
correct mathematical expressions. The former occurs through interactions of
growing children with human society, the latter through interactions of curious
scientists with nature. Just as a child discovers the correct words to use for
specific objects or ideas through social interactions, similarly a scientist learns
the appropriate expressions for a particular physical entity by interacting with
nature.
Homonymy: During the above process of infantile linguistic experimen-
tation, the growing child soon discovers that the same word can be applied
to very different physical structures. For example, the infant would discover
that the word “current”could denote either the flow of a river meandering
across the countryside, or the electricity streaming through a copper wire.
Another example it may come across could be the word “bow”, which could
connote either the front of a ship (as in “bow and stern”), a tied ribbon (as
in “bow-tie”), or a weapon to shoot projectiles with (as in “bow and arrow”).
This phenomenon is once again observed in scientific discovery as well.
During the initial stages of investigation, several different mathematical mod-
els and sets of terminology are applied to describe a certain physical entity or
empirical phenomena. The history of science deonstrates that, after a great
deal of time and effort, the vast majority of these models will be discarded
as each will be found to be failing in one area or the other. Newer and newer
models will be developed, each capable of explaining a wider spectrum of ob-
served physical properties or phenomena than the ones before. Ultimately,
physical reality will be most closely approximated by a finite group of par-
ticular and unique mathematical structures, each being able to describe a
certain set of empirically observed phenomena connected with the entity un-
der investigation. At this stage, the scientist can be seen to have discovered
the “proper” words and phrases to apply to a particular physical object or
empirical phenomenon.
The fact that several models may accurately describe a certain entity is
the direct counterpart of homonymy in linguistics. Nature is, of course, far too
complex to be reduced to one single set of equations. Hence, multiple models
14 Group Theory in Particle, Nuclear, and Hadron Physics

of varying degrees of accuracy and different areas of applicability exist and


are, indeed, to be expected.
Dialects: The specific mathematical models applied to a particular phys-
ical objet or empirical phenomenon are a direct parallel to the concept of a
dialect or jargon in linguistics. The former represents a subset of the language
of mathematics as applied to a particular physical entity; the latter represents
a subset of the language of a larger region as prevalent in a particular region
or social group. Hence, after the long process of experimentation is over, we
can say that the physicist/mathematician has finally developed the proper
dialect or jargon to apply to the object under consideration.
The entire purpose of science is then to continue to read nature through
this “exact” mathematical language, and develop various dialects to under-
stand each particular field or object of nature. The acquisition of a larger
vocabulary of this mathematical language leads to a greater fluency with na-
ture.
(c) Pure Mathematics versus Applied Mathematics
However, the above useful mathematics which is denominated as “applied”
constitutes a small part of the whole mathematical edifice that exits today. For
even a superficial analysis of this perceived “Language of Nature” reveals that,
at any one point of time, a substantial portion of this framework lies unused
by nature itself, finding no application in any field of science whatsoever.
This is the so-called “pure” mathematics [13]. It exists outside any physical
framework – a pure construct of human intellect, as many a mathematician
would have us believe.
In fact, it is a dream of many a mathematician to construct a mathe-
matics which can be labelled as “pure”, that is uncorrupted by any “lowly”
application to any concrete physical entity, that is “unsullied” through any
“degraded” utilization in the real material world. These individuals find a
thrill in creating something that is absolutely independent of any existing
material object.
Thus Hardy boasted [14], “I have never done anything “useful”. No dis-
covery of mine has made, or is likely to make, directly or indirectly, for good
or ill, the least difference to the amenity of the world.”
So obviously, though mathematics may be the language of nature (i.e., the
“applied” part), most of it is not (i.e., the “pure” part). What is that “pure”
mathematics? It seems to have a Platonic world of its own.
(d) Logical Positivism
The logical positivists attempted to comprehend this dichotomy by argu-
ing that knowledge possesses two distinct sources - the logical reasoning and
the empirical experience. According to them, logical reasoning leads to the
analytical a priori knowledge which embraces the field of pure mathematics.
In contrast, the empirical experience leads to the synthetic a posteriori knowl-
edge and this corresponds to the domain of applied mathematics. Explicitly
Basic Symmetry Concepts 15

or implicitly, such a “dichotomous” point of view of the intrinsic structure


of mathematics is held by many scientists, mathematicians, and philosophers
today.
And if mathematics is indeed the language of nature, it must be explained
why these two mutually exclusive categories of language exist. How come
there is a smaller reservoir of language which nature communicates with (i.e.,
applied mathematics) while there is a larger reservoir of unused language
(i.e., pure mathematics)? How and why does this unused language (“pure
mathematics”) come into existence?
Gibberish: Here, the field of linguistics once again offers a possible paral-
lel. Within the periphery of a particular language, any sound which does not
fall within the specified category is defined as gibberish.
Now, for each acceptable sound in any given language there are clearly
many more gibberish sounds outside that language. Thus, the range of gib-
berish sounds outside any language is much larger than that of acceptable
phonemes in the language. This is the same as in the case of “pure” mathe-
matics.
On the basis of what has been stated so far, “pure” mathematics should
therefore be understood so as to belong to the honourable category of “gib-
berish”. No offense is meant by this statement, but as far as the language
of nature is concerned, if the relevant applied mathematics is the exact and
accurate vocabulary of nature, then the “pure” mathematics must necessar-
ily be treated as “gibberish” in the framework of what one understands as a
“language”.
For, just as a child can produce a large number of gibberish sounds in
ordinary language, so can a mathematician produce a vast amount of “gib-
berish” mathematics. Just as the structure of physical vocal biology allows
us to create a large amount of gibberish sounds, so also the mathematical
reality of nature seems to be structured in such a manner that it permits the
construction of an enormous amount of mathematical “gibberish”.
(e) New Mathematics
Sometimes scientists would have to develop ab initio the necessary mathe-
matics to understand physical reality. For example, to understand the empir-
ically determined Kepler’s laws of planetary motion, Newton had to develop
the requisite calculus to do so. The very fact that Newton was actually able
to acquire the necessary mathematical vocabulary made it possible for him to
appreciate the effects of gravity. The physical “book” of gravity was “read”
only because the necessary mathematics could be simultaneously developed. It
was to “read” the other physical effects as well, that Newton’s contemporary,
Leibniz was independently developing the required mathematics of calculus.
Hence the requisite mathematical language of calculus was basic and essential
to an understanding of gravity and dynamics in physical nature. Simply put,
had it not been possible to develop the language of calculus, one would not
have been able to read nature any better.
16 Group Theory in Particle, Nuclear, and Hadron Physics

Now, the basic mathematics of calculus could be developed by scientists


themselves (i.e., Newton and Leibniz) as fortunately it did not necessarily
require a too sophisticated pre-existing mathematical framework. Their work
was simplified by the fact that the foundations of calculus had already been
laid by earlier mathematicians. It was not just for nothing that Newton had
stated that he had risen on the shoulders of giants.
Very often development in science is hampered if the adequate and appro-
priate mathematical framework does not exist. Therefore, had the algebra of
tensors not been developed by Einstein’s and Hilbert’s contemporary math-
ematicians, they would never have been able to arrive at their equation of
motion in General Theory of Relativity in 1915.
(f ) Transformation from Pure to Applied Mathematics
The hoary history of science is replete with instances of “pure” mathe-
matics finding valuable applications within the real physical world and thus
becoming “applied” mathematics.
One example relates to Hardy who, as we have described earlier, was ex-
ceedingly proud of the “purity” of his mathematics. In fact, his work on infinite
series in number theory was originally deemed to be a perfect example of this
“pure” and “unsullied” mathematics. Today, this framework has been very
“sullied” indeed, finding widespread application in the domain of cryptogra-
phy and communications theory.
Likewise, the theory of mathematical matrices were once thought to be
beyond any applications in nature. Today, these form the bread and butter of
quantum physicists.
Hence, what may be mathematical “gibberish” today may turn out to
attain the status of a proper and correct set of words in the language of
nature to describe physical entities as relevant “applied” mathematics.
Note that we are not using the word “gibberish” in any disrespectful man-
ner. It is used here as what it really means within what we mean by a language.
At present so many bizzare combinations of alphabets are finding acceptability
within SMS messages and texting. So what is called gibberish today may be-
come meaningful tomorrow. What is so denominated pure mathematics today
may become applicable mathematics tomorrow.
Now, Margolis [15] has pointed out that mathematical language is context
free. But the spoken language depends on the context of its usage. Note that
a gibberish word is definitely context free. It is, in a manner, invented for
its own sake. It exists as its creater wanted it that way. This freedom from
contexuality is what pure mathematics is all about. But once applied, it gains
a context to work within.
As an example, the special unitary group in two dimensions SU(2) (as
will be demonstrated) was a pure mathematical structure until the early nine-
teenth century. To explain the fine stucture of the hydrogen atom, Uhlenbeck
and Goudsmit postulated the existence of two-valued spin angular momen-
tum. And the above SU(2) was the natural structure to provide the relevant
Basic Symmetry Concepts 17

language for this new property [16]. Subsequently, in 1934 Heisenberg pro-
posed that the same mathematical structure of SU(2) could also provide the
correct langauge to describe how (proton, neutron) be treated as a doublet
of what we now call the isopin-SU(2) group. In recent years, yet another in-
dependent interpretation of the language of SU(2) has been proposed. This
is to treat the (He3 , H 3 ) pair as a doublet of a new SU (2)A , the nusospin
group. So the same once-pure mathematical structure of SU(2) is providing
new “words” for the language of nature.
(g) Brain Evolution
Extending the above application of linguistics to the language of nature,
we find that the domain of cognitive linguistics provides certain fascinating
insights into the historical development of mathematics. For it appears that
normal human languages and the language of nature are functionally and neu-
roanatomically independent, and are probably processed in entirely different
parts of the brain. This is indicated in a recent study in which Rosemary
Varley et al. [17] analyzed three men with brain damage which severely de-
graded their ability to handle normal grammar. An important observation of
the research group was that these men, while having lost the ability to pro-
duce or comprehend grammatical syntax, retained the full ability to perform
complex numerical computations using Arabic numerals. In fact, the individu-
als demonstrated proficiency in mathematical syntax, including the ability to
handle structure-dependent conceptions such as brackets and recursion. This
is a clear indication that the ability to handle the language of mathematics is
completely independent of the capacity to utilize human languages, and that
these two distinct skill-sets register differently in the human brain, perhaps
even requiring completely separate regions of the cerebrum.
Another factor supporting this thesis comes from the exactitude of the
languages themselves. For the set of human languages such as English, Arabic
and Persian are imprecise and fuzzy, while the language of nature is precise
and exact. Indeed, it would have been rather puzzling if these two distinct sets
of languages were to register and be controlled by the same neuroanatomical
region of the cerebrum.
The findings of Varley et al. cited above provide convincing evidence that
linguistic grammar and mathematical reasoning are processed in different
manners in the human brain, possibly registering in different regions of the
cerebrum itself. It is hence possible that these distinct capabilities arose at
different stages during the process of human evolution. Furthermore, they
may be activated at different periods of infant growth, the children acquiring
these capabilities at different periods of life, due to different requirements of
adaptation needed for them to become essential for survival.
The first, the spoken language, arose as a result of social interaction and
as a result of demands of survival for food, etc. while the second, the language
of mathematics, arose as a result of man’s interaction with nature. As man
spends more time with nature than with other human beings, the second
18 Group Theory in Particle, Nuclear, and Hadron Physics

language must be more naturally acquired than the first one. This point is
also supported by the fact that other creatures have been interacting with
nature for a longer period of time than what we humans have been doing. Do
they have a language of mathematics? Indeed they do. When a bird needing
to feed four chicks in its nest actually brings back four insects to feed them,
then indeed it has acquired the rudiments of mathematics. Hence it is clear
that humans in the course of evolution must have learned elements of the
language of nature well before they learned to speak. Therefore, it may come
as a surprise to some, but mathematics as a language of nature, albeit in a
more elementary form, must have been available to species other than Homo
sapiens. Indeed current research shows that acquisition of spoken language
may be a much later development in human evolution. In fact, the growth of
the human brain and the faculty of (spoken) language acquisition may have
been simultaneous [18].

1.6 Solutions of Problems


Problem 1.1:
We have seen that for the 3-gon figure, the cyclic group C3 has 3 elements.
Given threefold axis of reflection the dimension of D3 is 6. For 5-gon there
are five elements of rotation around its axis of symmetry plus five reflections
to give 10 elements for the group D5 . Generalize to N-gon, the cyclic group
CN has N-elements of rotations around its symmetry axis and N-reflection
symmentries to give 2N elements of the dihderal group DN .
————————————————–
Problem 1.2:
Because CN forms a subgroup of DN and so identity should be there.
Reflections do not form a subgroup; they are not closed on themselves.
————————————————–
Problem 1.3:
With Lagrangian L(q, q̇) the Lagrange equation of motion is
d ∂L ∂L
( )− =0
dt ∂ q̇ ∂q
Unmeasurability of absolute value of q means that ∂L ∂L
∂q = 0 and so ∂ q̇ =
constant. Now as ∂L∂ q̇ =p, the generalized momentum is conserved. Taking q
as the position coordinate we find that the linear momentum is conserved.
If we took q to be the angular coordinate, then this shows that the angular
momentum is conserved. Hence both the homogeneity and the isotropy of
space follow from the same Lagrange equation of motion.
————————————————–
Chapter 2
Group Theory

2.1 Definitions and Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19


2.2 Subgroups and Cyclic Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.3 Cosets and Normal Subgroups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.4 Factor Group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.5 Homomorphism and Isomorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
2.6 Torsion Group and Betti Number . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
2.7 Appendix B: Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
2.8 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

2.1 Definitions and Examples


We have seen how in a particular symmetry, say C3 , a rotation by 2 π3
around its symmetry axis leaves the equivalent triangle invariant. Successive
rotations by 2 π3 likewise leave the system invariant. There are only three
independent transformations, 1, 2 π3 and 4 π3 rotations. The next rotation to
6 π3 = 2π brings the shape back to itself, i.e., the identity 1. Thus the trans-
formations here form a closed set. The same holds for a more general CN and
DN .
Let us now generalize to such a symmetry as those particular transfor-
mations which exhibit the property that two successive transformations of a
particular type are equivalent to a single transformation of the same type.
Symbolically say first,
0
x=x +a

Next
0
x00 = x + b

Then

x00 = x + (a + b) = x + c

Such transformations exhibiting such a property are denominated a group


if, in addition, they follow certain conditions.

19
20 Group Theory in Particle, Nuclear, and Hadron Physics

A group is a set of elements G = {a, b, c, ....} obeying a single law of com-


position (defined as a • b) which must satisfy the following constraints:

(1) Closure: a, b ∈ G then a • b = c ∈ G

(2) Associativity: a, b, c ∈ G and (a • b) • c = a • (b • c)

(3) Identity: These exists i ∈ G such that for a ∈ G i • a = a • i = a

(4) Inverse: For every a ∈ G these exists an inverse a−1 ∈ G such that
a • a−1 = a−1 • a = i
(Note: The symbol • denotes any mathematical composition such as
+, −, ×, ÷ and any other mathematical operation as we see below)
If in addition a • b = b • a then the group is termed Abelian. If a • b 6= b • a
then the group is called non-Abelian. Note that a group is comprised of two
entities; a set G and a binary operation • on G. Thus there are two ingredients
involved. When the existence of a group G is asserted, the presence of an
associated binary operation is implicitly implied [19].

————————————————–
Problem 2.1: Prove that the inverse of an element in a group is unique.
————————————————–

Example (a):
Z is a set of all integers. Now,

Z = {o, ±1, ±2, ±3, ......} (2.1)


Does it form a group under the composition law (binary operation) of ad-
dition (so here • ≡ +)?

(1) Closure: ∀a, b ∈ Z; a + b = c∈ Z ; satisfied.

(2) Associativity: (a + b) + c = a + (b + c), ∀a, b, c ∈ Z; satisfied.

(3) Identity: 0 + a = a = a + 0 so additive identity is 0; satisfied

(4) Inverse: −a ∈ Z for a ∈ Z; −a + a = 0 = a − a ; satisfied.

Hence Z is a group under addition. Also as a + b = b + a, it is Abelian. It


is infinite dimensional, as the number of elements is infinite.

————————————————–
Problem 2.2: Is Z a group under the binary operation of multiplication?
————————————————–
Group Theory 21

Example (b): Is G, a set of even integers G = {0, ±2, ±4, ......}, a group
under the binary operation of addition?

(1) Closure: Sum of even integers is even; satisfied.

(2) Associativity: (a + b) + c = a + (b + c); a, b, c ∈ G; satisfied

(3) Identity: 0 identity; satisfied.

(4) Inverse: For a ∈ G, a−1 = −a as (−a) + a = a + (−a); satisfied.

Hence, G is an Abelian group.


0 0
Example (c): Is G as a set of odd integers G = {0, ±1, ±3, ±5......} a
group under the binary operation of addition?
0 0 0
(1) Closure: For a, b ∈ G , a + b = c ∈
/ G (e.g., 3 + 3 = 6 ∈
/ G ) so it is not
a group under addition.

Example (d):
Is the set of rational numbers {Q = 0, ab ; a, b ∈ Z and b 6= 0} a group under
addition?
a1 a2 a1 b2 +a2 b1
(1) Closure : b1 + b2 = b1 b2 = ab ∈ Q; satisfied.
   
a1
(2) Associativity : b1 + ab22 + ab33 = ab11 + ab22 + ab33 ; satisfied.

a a a
(3) Identity : 0 is identity ; b +0=0+ b = b ∈ Q ; satisfied.

(4) Inverse : a
b ∈ Q, a−1 = − ab ∈ Q ; a
b − a
b = 0; satisfied.

So it is an Abelian group.

————————————————–
Problem 2.3: Is the set Q0 of rational numbers without 0, a group under
multiplication? Is it a group with 0 included (full Q)?
————————————————–
Problem 2.4: Demonstrate that the set C of all complex numbers forms
an infinite Abelian group with respect to addition.

C = {a1 + ib; a, b ∈ R, i2 = −1}

R is the set of real numbers.

————————————————–
22 Group Theory in Particle, Nuclear, and Hadron Physics
TABLE 2.1:
Composition table
of the cube roots of
unity
1 ω ω2
1 1 ω ω2
ω ω ω2 1
ω2 ω2 1 ω

Note: nth roots of unity are given as


1 1
xn = 1 → x = (1) n = {cos(2πr + 0) + isin(2πr + 0)} n (2.2)
So
i
x = e2πr n ; r = 0, 1, 2, · · · (n − 1) (2.3)
Hence the cube roots of unity are
i i
1, e2πr 3 , e4πr 3 (2.4)

2π 3i
With e = − 21 + i 23 = ω and
i

e4π 3 = − 12 − i 23 = ω2

Note ω · ω = ω 2 = 1
ω = ω −1 as ω 3 = 1

Note as a set {1, ω, ω 2 }, the cube roots of unity do form a group under
w2 w2
multiplication. Taking 1 as identity w−1 = w1 = w.w 2 = 1 is satisfied. It
is clearly Abelian and of course is finite-dimensional (as three numbers of
elements exist).
We can express the entire group structure of a finite-dimensional group in
a composition table. For the cube roots of unity, the composition table is
given in Table 2.1.
Here the first column on the left is the first operation (i.e., acting on the
right), the first row is the second operation (on the left). So in say A, B ∈
(some group), in AB the operation B is along the column and A is along the
row. This distinction is significant for a non-Abelian group (we will see in
the following examples). However, for an Abelian group like the one above,
the order, left or right, does not matter. Now composition tables are a useful
graphical manner of displaying the structure of the group wherein the number
of group elements (the order of the group) is small.

————————————————–
Group Theory 23

(a) (b)
3 3

1, lt

1 2 1 2
I.

(C)

<f.
4
3

I;

1
2 <f,
I,

FIGURE 2.1: Symmetries of an equilateral triangle and a square

Problem 2.5: Show that the two square roots and the four fourth roots
of unity form a multiplicative group. Write the composition table of these.
————————————————–
Problem 2.6: If Q+ is the set of all positive rational numbers ( ab with the
composition defined as a • b = ab
2 , does this composition operation provide a
group structure for Q+ ?
————————————————–

Now let us look at the cyclic group C3 . Its elements are obtained as
1, 2 π3 , 4 π3 rotations. We label the vertices of an equilateral triangle by in-
dices 1,2,3 as in Figure 2.1(a,b).
Then 1 means no change, 2 π3 rotation denotes (counterclockwise) rotation
around its symmetry axis through the centre and perpendicular to the plane
of the paper; this is 1 → 2 → 3 → 1 rotations. Next 4 π3 rotations connotes
1 → 3 → 2 → 1. We denote these elements as ρ0 , ρ1 and ρ2 , respectively. (Note
for CN we shall denote ρi as similar rotations). Thus, the composition table
of C3 is given in Table 2.2.
Next let us study D3 , the dihedral group. It is built up of rotations as in
C3 plus reflections – geometrically corresponding to the turning over of the
triangle about an axis li (see Figure 2.1(b)). Thus µ1 keeps 1 stationary and
flips over 2↔ 3. Similarly, µ2 keeps 2 fixed and flips over 3 ↔ 1 while µ3 leaves
3 fixed and 1 ↔ 2. Thus, µ2 ρ1 = µ3 (where ρ1 is first operation and µ2 the
second operation as per our convention). We elucidate these operations next
24 Group Theory in Particle, Nuclear, and Hadron Physics
TABLE 2.2:
Composition table
of the cyclic group
C3
ρ0 ρ1 ρ2
ρ0 ρ0 ρ1 ρ2
ρ1 ρ1 ρ2 ρ0
ρ2 ρ2 ρ0 ρ1

TABLE 2.3: Composition


table for the group D3
ρ0 ρ1 ρ2 µ1 µ2 µ3
ρ0 ρ0 ρ1 ρ2 µ1 µ2 µ3
ρ1 ρ1 ρ2 ρ0 µ2 µ3 µ1
ρ2 ρ2 ρ0 ρ1 µ3 µ1 µ2
µ1 µ1 µ3 µ2 ρ0 ρ2 ρ1
µ2 µ2 µ1 µ3 ρ1 ρ0 ρ2
µ3 µ3 µ2 µ1 ρ2 ρ1 ρ0

(note that here it is not a matrix operation, but a schematic representation


of the sequence of rotations of 1,2,3),
   
3 2
ρ1 =
1 2 3 1

       
3 2 3 3
µ2 ρ1 = µ2 = = µ3
1 2 3 1 2 1 1 2

Also

       
3 3 2 3
ρ2 µ3 = ρ2 = = µ1
1 2 2 1 1 3 1 2

And so ρ2 µ3 = µ1 . The complete composition table for D3 is given in


Table 2.3. Note that the group D3 is a non-Abelian group of six elements.
Note: 1. The first box of ρ-terms is exactly the same as that of C3 (see
Table 2.2) which is a complete group in itself. Hence we denote C3 ⊂ D3 , i.e.,
C3 is a subgroup (which itself is a group) of D3 .
2. If we replace (ρ0 , ρ1 , ρ3 ) by (1, ω, ω 2 ), respectively, then the C3 table
is identical to the table of the cube roots of unity. The sameness of the two-
composition tables is termed isomorphism. We will talk more about subgroups
and isomorphism next.

————————————————–
Group Theory 25
TABLE 2.4: Reflections and
rotations as blocks in the D3 group
ρ − term µ − term
ρ − term ρ − term µ − term
µ − term µ − term ρ − term

Problem 2.7: Verify that µ3 ρ1 = µ1 ; µ1 ρ1 = µ2 ,µ1 µ3 = ρ2


————————————————–

Note that ρi is for rotations and µi is for mirror images in bisectors of


angles. Further note the composition table for D3 is broken into four blocks
as shown. Algebraically these four blocks give a “group” of order 2. We provide
this blocking of the D3 composition table (Table 2.4).
We interpret this table geometrically. It implies that the product of ro-
tations (ρ-terms) is a rotation, the product of a rotation and a reflection
(µ − terms) is a reflection, and the product of two reflections is a rotation.
This splitting up of a group into blocks which form a group by themselves
shall be studied shortly. This mathematical intricacy shall have deep physical
significances.
Next consider D4 the 4th dihedral group (also called the octic group) of
symmetries of a square given in Figure 2.1(c). Note that D4 is a non-Abelian
group of 8 elements [20] ,[21].
Whereas ρi are rotations (of 0, π2 , π, 3π
2 ) as (ρ0 , ρ1 , ρ2 , ρ3 ), respectively, µi
are mirror images in perpendicular bisectors of sides given by l1 and l2 , and
δi for diagonal flips around the d1 and d2 axes.
       
4 3 2 1 1 2 4 3
µ1 ρ2 =µ1 = = µ2
1 2 3 4 4 3 1 2
       
4 3 2 1 2 3 4 3
δ 1 ρ2 =δ1 = = δ2
1 2 3 4 1 4 1 2

So µ1 ρ2 = µ2 and δ1 ρ2 = δ2
The complete composition table of D4 is given in Table 2.5.
Note that the first box is indicating the group structure of the cyclic rota-
tion group C4 . As such it is a subgroup of D4 , C4 ⊂ D4 . D4 is a larger group
containing a small group C4 within it.

————————————————–
Unsolved Problem 2.1: Confirm the products implied in the table for
D4 .
————————————————–

Next, we define “Residue Classes Modulo m”. If a1 , a2 ∈ Z then a1 =


26 Group Theory in Particle, Nuclear, and Hadron Physics
TABLE 2.5: Composition table of D4
ρ0 ρ1 ρ2 ρ3 µ1 µ2 δ1 δ2
ρ0 ρ0 ρ1 ρ2 ρ3 µ1 µ2 δ1 δ2
ρ1 ρ1 ρ2 ρ3 ρ0 δ2 δ1 µ1 µ2
ρ2 ρ2 ρ3 ρ0 ρ1 µ2 µ1 δ2 δ1
ρ3 ρ3 ρ0 ρ1 ρ2 δ1 δ2 µ2 µ1
µ2 µ1 δ1 µ2 δ2 ρ0 ρ2 ρ1 ρ3
µ2 µ2 δ2 µ1 δ1 ρ2 ρ0 ρ3 ρ1
δ1 δ1 µ2 δ2 µ1 ρ3 ρ1 ρ0 ρ2
δ2 δ2 µ1 δ1 µ2 ρ1 ρ3 ρ2 ρ0

a2 (mod m) if (a1 − a2 ) is divided by m. a1 = a2 (mod m) is read as “a1


congruent to a2 modulo m”. Or take it as when a1 is divided by m then a2
is the remainder. Any integer a is congruent (mod m) to its remainder when
divided by m and there are m possible remainders ( 0, 1, 2, 3, 4, · · · m−1). Next
(m, m + 1, m + 2, ·2m − 1) has remainders again as (0, 1, 2, 3, 4, ......m − 1)
and so on. So all the integers are separated into classes. That is, all integers
which are congruent modulo m are put in the same class.
So in the above m disjoint classes are called residue classes modulo m.
Thus, a1 = a2 (mod m) classifies all the integers into ({0}, {1}, {2}, · · · {m−1}),
i.e., m classes

Example: {0}, {1}, {2}.{3}, {4} are residue classes modulo 5. Sum of
residue classes,

{0} + {3} = {3} as 0 + 3 = 3 < 5

{1} + {3} = {4} as 1 + 3 = 4 < 5

{2} + {3} = {0} as 2 + 3 = 5 = 1 × 5 + 0 = 0 mod(5)

{3} + {4} = {2} as 3 + 4 = 7 = 1 × 5 + 2 = 2 mod(5)

Thus, in general if {r1 } and {r2 } are two residue classes of modulo m then
under addition,

{r1 } + {r2 } = {r1 + r2 } if r1 + r2 < m


= {k} if r1 + r2 ≥ m (2.5)
Where k is the least non-negative remainder when r1 + r2 is divided by m.
Now we demonstrate that the residue classes modulo m forms a group
with respect to residue classes (i.e., addition modulo m):
So the members of the set in the group are
Group Theory 27
TABLE 2.6:
Composition
table of addition
modulo 4
0 1 2 3
0 0 1 2 3
1 1 2 3 0
2 2 3 0 1
3 3 0 1 2

G = [{0}, {1}, {2}, .....{p}, {q}, .......{m − 1}]

(1) Closure: {p} + {q} = {p + q} if p + q < m ∈ G

={k} is the lowest remainder when (p + q) is divided by m

(0 ≤ k ≤ m) ∴ {k} ∈ G; Satisfied.

(2) Associativity: ({a} + {b}) + {c} = {a + b} + {c} = {(a + b) + c}

= {a + (b + c)} = {a} + ({b} + {c}); Satisfied.

(3) Identity: Since {0} + {r} = {r} 0 ≤ r ≤ m

So the identify for addition is {0}, i.e., zero-class; Satisfied.

(4) Inverse: additive inverse for {r} is {m − r}

{m − r} + {r} = {m} = {0}; Satisfied.

Hence this is a group. In addition:

{p} + {q} = {p + q} = {q + p} = {q} + {p}. So it is an Abelian group.

The above group is denominated Zm .


We provide the composition table of the group Z4 (addition modulo 4) in
Table 2.6.
Note that there is only one independent group of order 2 (as given in
Problem 2.5) and of order 3 as given in Table 2.1 and Table 2.2, which are
actually the same as (1, ω, ω 2 ) ↔ (ρ0 , ρ1 ρ2 ), i.e., isomorphic to each other). We
saw Z4 is a group of order 4. It turns out that there is one more independent
group of order 4, it is termed the 4-group or given by the symbol V (V is an
abbreviation for the German word Viergruppe). It is given in Table 2.7. V has
four members given as (e, a, b, c), when e is the identity.
28 Group Theory in Particle, Nuclear, and Hadron Physics
TABLE 2.7:
Viergruppe V or
4-group
e a b c
e e a b c
a a e c b
b b c e a
c c b a e

————————————————–
Problem 2.8: Draw the composition table of Z6 (addition modulo6) with
six elements.
————————————————–
Problem 2.9: Using the logic of sameness that was used between the
group of cube-root of unity and C3 , establish how Z4 and the fourth roots of
unity are the same (i.e., how these are isomorphic to each other).
————————————————–

2.2 Subgroups and Cyclic Groups


A subset H of a group G is a subgroup of G if and only if

1. The binary operation of G is closed on H.

2. The identity e of G is in H.

3. For all a ∈ H it is true that a−1 ∈ H also.

Note that part (1) of the above is obvious from the definition of the sub-
group itself.

Part (2): Let e and e0 be identities of G and H, respectively,

a ∈ H → e0 a = a

a ∈ G → ea = a

and so in G, ea = e0 a → e = e0 (using right cancellation).

Thus, e ∈ H
Group Theory 29

Part (3): ax = e must have a solution in H.


So x = a−1 in H. Hence the inverse a−1 of a in G is also the inverse of a
in H.
Another useful way of defining a subgroup is through H ⊂ G. This is a
subgroup of G if for each pair of elements h, h0 ∈ H, we also have h−1 h0 ∈ H.

1. Closure: Satisfied from definition

2. Associativity: H⊂ G so satisfied.

3. Identity: since for h = h0 h−1 h0 = h−1 h = e ∈ H, satisfied.

4. Inverse: For h0 = e h−1 h0 = h−1 e = h−1 ∈ H. Also if h =


−1 −1 −1 0 0
h , (h ) h = hh ∈ H; Satisfied.

Hence, H is a subgroup of G.
We have already seen the subgroup structure in C3 ⊂ D3 . Note that the
identity of C3 is ρ0 , which is also an identity of the larger group D3 . The rest
of the numbers of D3 (µ1 , µ2 , µ3 ) do not for a subgroup simply because the
identity does not exist in this subset, having been usurped by C3 .
The group G may be considered to be a subgroup of itself, but it is de-
nominated a trivial subgroup or improper subgroup. The identity e is
also a subgroup but again is called trivial. So each group possesses two triv-
ial subgroups – the full group itself and e the identity on its own. All other
subgroups (if they exist) are termed proper subgroups.
Next look at the Z4 group in Table 2.6. Note that the subset of (0, 2) forms
a subgroup as 0 + 0 = 0, 2 + 0 = 2, 2 + 2 = 4, which are all belonging to the set
(0, 1, 2, 3) of Z4 and are closed under the same, addition modulo4 operation.
Note that, e.g., the subset (0, 3) is not a subgroup as 3 + 3 = 6 = 2(mod 4)
and 2 ∈ / (0, 3).
Next the group V has three proper subgroups as the set (e, a), (e, b), (e, c).
However the subset (e, a, b) is not a subgroup as ab = c ∈ / (e, a, b). Note that
(e) is a subgroup of these two element subgroup of V .
We display the complete subgroup structure of a group as a lattice diagram.
First row is the original group G. The next group lists all the subgroups of G.
The row below it lists all the subsequent subgroups of the above row. And so
on. The lattice diagram for the subgroup structure of Z4 V is given in Figure
2.2(a).

————————————————–
Problem 2.10: For the octic group or D4, using Table 1.5 draw the lattice
diagram indicating the subgroup structures.
————————————————–
30 Group Theory in Particle, Nuclear, and Hadron Physics
(a)

Z
4 V

{0,2} *e,a> *e b> {e,c}

<0> <e>

(b)
D3

*po'px'pa> { Po'^> {PO ' ^ > < P0 ' ^3 :

<a"„>

FIGURE 2.2: Lattice diagram of the subgroups of the groups Z4 , V and D3

We also give the lattice diagram for the group D3 – the dihedral group in
three dimensions, in Figure 2.2(b).
Now in the cube roots of unity, the elements are (1, ω, ω 2 ). All the elements
are generated by w : as w2 and w3 = 1. One can continue with this cycle
and thus it is a cyclic group. Furthermore, C3 with (ρ0 , ρ1 , ρ2 ) is identically
generated by the element ρ0 . Thus, we define a cyclic group as one where
all the elements are generated by one particular element which is termed the
generating element, with all its distinct powers. Higher powers do not yield
any new element and thus the identity element arises from the generating
element raised to a power equal to the order of the cyclic group.

G = {a, a2 , a3 , · · · an = e} order (G) = n (2.6)


th
Cn is the simplest cyclic group and so are the n roots of unity. All cyclic
groups of the same order are basically identical (isosmorphic).

————————————————–
Problem 2.11: Show that the 4th roots of unity are a cyclic group. What
is its generating element?
————————————————–

A subgroup of a cyclic group is cyclic itself. We denote the generating


element (generator) as < a >. We saw that {0, 2} ∈ Z4 is a subgroup. It is
cyclic as 2 + 2 = 0 so < 2 > of Z4 and also < 1 > generates all the elements
Group Theory 31

of Z4 {0, 1, 2, 3}. < 3 > also generates the whole group Z4 as 3, 3 + 3 = 2,


3 + 3 + 3 = 1, 3 + 3 + 3 + 3 = 0 so ¡3¿=G. However, V is not cyclic. But
its subgroup < a >= {e, a}, < h >= {e, b}, < c >= {e, c}. So its proper
subgroups are cyclic.

————————————————–
Problem 2.12: Demonstrate that Z under addition is a cyclic group.
What are its generators?
————————————————–

Let us state a

Theorem: An element of ap of finite cyclic group G of order n is a gener-


ator of G iff (n, p) = 1, i.e., p is prime to n and 0 < p < n.

————————————————–
Problem 2.13: Prove the above theorem.
————————————————–

Example: Given a cyclic group as order 8, how many elements of it are


generators of the group?
We have (a, a2 , · · · a8 = e) so < a > is a generator.
ap a prime. Then (p, n) = 1 or (p, 8) = 1 → p = 3, 5, 7. Also as a−1 = a7
the total number of generators of this 8 – order cyclic group is:
< a >, < a3 >, < a5 >, < a7 >.
Next consider the group Z under addition.
First, < n >= 3, 3 + 3 = 6, 3 + 3 + 3 = 9 etc. and 0; -3, −3 + (−3) =
−6,−3 + (−3) + (−3) = −9.
So the cyclic subgroup generated by 3 consists of all the multiples of 3:
positive, negative and zero. This is 3Z or < 3 >. This is 3Z or < 3Z >. So
nZ is the cyclic subgroup < n > of Z.

————————————————–
Problem 2.14: Discuss the cyclic group {a, a2 , a3 , a4 , a5 = e}, its gener-
ators and its subgroup.
————————————————–
32 Group Theory in Particle, Nuclear, and Hadron Physics

2.3 Cosets and Normal Subgroups


Cosets of subgroups are useful as these divide a group into disjoint sets.
Let us state,

Lagrange theorem:: If H is a subgroup of G of order o H (order of G is


O
O G
G), then O H = an integer.

So the order of G must be an integral multiple of the order of H for all


H ⊂ G.
O O
G
For the trivial subgroup: (i) {e} the O H = 1 and so O H = 1G =O G
O
G
integer. (ii) H = G so O H = O G and so O H =1 and thus the Lagrange
theorem is automatically satisfied for these.
Now for the proper subgroup H, as H is a proper subgroup it means that
its order is less than that of G and so there is an element 0 a0 which belongs to
G and not to H.
Left cosets are defined by multiplying all elements of H on the left by 0 a0
and collecting all these products into a set defined by aH.

aH = {a, ah1 , ah2 , ....} = {ahi | hi ∈ H and a ∈


/ H} (2.7)
Similarly one defines the right cosets of H by Ha and in general these
will not to be the same as the left coset.
Example: Let us determine the left coset of 3Z as a subgroup of Z under
addition.
3Z = 0 + 3Z, is itself the left coset. So are (1 + 3Z) and (2 + 3Z).
So the reminder of any integer divided by 3 is an integer 0 ≤ r ≤ 3. Hence
the above are the left cosets.

————————————————–
Problem 2.15: Demonstrate that Z6 splits up as left coset of the subgroup
of H = {0, 3}
————————————————–

Here we list a few important properties of cosets. Note that here we select
the left cosets (but the same would hold for the right cosets also).
(a). The elements of aH are distinct. If not, then ahi = ahj → hi = hj .
Thus aH has ordered o H of distinct elements, the same as in H itself.
(b) aH is not a group. To be a group e (identity) should be contained in
it. So for hi ∈ H, ahi = e → a = h−1i ∈ H. So a should be an element of H
to render it a subgroup. But this violates the condition that a ∈/ H.
(c) (aH) ∩ (Ha) 6= 0. Both aH and Ha contain the element a (i.e., ae = a
in aH and ea = a in Ha).
Group Theory 33

(d) (aH) ∩ (H) = 0. If there is a single common elements aHi = hi, then
a = hi hi−1 ∈ H which is against the very definition itself. And the elements
of (aH) are different from those of H.
(e) Two left cosets aH and bH are either disjoint are identical. Hence if
ahi = bhi then a = bhi h−1 i = bhi ∈ bH and a ∈ bH. Next for any element
ahl ∈ aH, ahl = bhj h−1l = bhk ∈ bH. So these are identical or (aH)∩(bH) = 0
Hence as per point (e) above, H and disjoint cosets of H divide or partition
the group G into disjoint sets. This division of G into disjoint sets is termed
G
the left quotient set H .
Note that in the above discussion each coset contains O H elements and
the quotient set spans the whole group G, thus the number of sets in the
quotient set multiplied by O H = O G. These are integers and this is what the
Lagrange Theorem states. The number of sets in the quotient set, i.e., the
number of distinct cosets aH plus H itself, is called the index of H in G.
Clearly this index is independent of left cosets or right cosets.
If the left and right cosets for a particular subgroup are the same, i.e.,

sH = Hs f or all s ∈ G (2.8)
Then H is said to be the normal subgroup (it is also denominated the
invariant subgroup)
Example A
Given the Viergruppe V in Table 2.7 let us determine the left and right
cosets of the proper subgroups given in the lattice diagram in Figure 2.2. So
V = {e, a, b, c}. Note that one can always rearrange the elements in a group
or a subgroup. So we can write V = {c, a, e, b}.

V1 = {e, a}, bV1 = {b, c}, cV1 = {c, b} = bV1

V2 = {e, b}, aV2 = {a, c}, cV2 = {c, a} = aV2

V3 = {e, c}, aV3 = {a, b}, bV3 = {b, a} = aV3

These are left coset. Now right coset.

V1 = {e, a}, V1 b = {b, c}, V1 c = {c, b} = V1 b

V2 = {e, b}, V2 a = {a, c}, V2 c = {c, a} = V2 a

V3 = {e, c}, V3 a = {a, b}, V3 b = {b, a} = V3 a

Notice that for V the left and right coset are identical and so that all
subgroups of V are normal subgroups. There are three normal subgroups
here.
34 Group Theory in Particle, Nuclear, and Hadron Physics

————————————————–
Problem 2.16: For an Abelian group prove that all the subgroups are
normal.
————————————————–

Example B: For the group D3 with elements {ρ0 , ρ1 , ρ2 , µ1 , µ2 , µ3 }, the


subgroup structure is illustrated in the lattice diagram of Figure 2.2(b).
So there are three order-two subgroups and one of order three.

H1 = {ρ0 , µ1 }, H2 = {ρ0 , µ2 }, H3 = {ρ0 , µ3 }, H4 = {ρ0 , ρ1 , ρ2 }

Take the left and right cosets of H3 = {ρ0 , µ3 }

µ2 H3 = {µ2 ρ0 , µ2 µ3 } ; H3 µ2 = {ρ0 µ2 , µ3 µ2 }

= {µ2 , ρ1 } = {µ2 , ρ2 }

ρ2 H3 = {ρ2 ρ0 , ρ2 µ3 } ; H3 ρ1 = {ρ0 ρ1 , µ3 ρ1 }

= {ρ2 , µ1 } = {ρ1 , µ1 }

µ1 H3 = {µ1 , ρ2 } = ρ2 H1 ; H1 µ1 = {µ1 , ρ1 } = H1 ρ2

So the left and the right cosets of H3 = {ρ0 , ρ1 } are not the same so H3 is
not a normal subgroup.

For H1 = {ρ0 , µ1 } left cosets are ρ1 H1 = {ρ1 , µ3 } and ρ2 H1 = {ρ2 , µ3 }.

That is D3 is divided into 3 disjoint sets ({ρ1 , µ1 } | {ρ1 , µ2 } | {ρ2 , µ3 }). It


can easily be verified that H1 is not normal and hence the same for H2 also.
So all the proper order-two subgroups of D3 are not normal.

However H4 = {ρ0 , ρ1 , ρ2 }

µ3 H4 = {µ3 ρ0 , µ3 ρ1 , µ3 ρ2 } = {µ3 , µ1 , µ2 } = µ2 H4 = µ1 H4

Next,

H4 µ3 = {ρ0 µ3 , ρ1 µ3 , ρ2 µ3 } = {µ3 , µ2 , µ1 } = H4 µ2 = H4 µ1

Thus, the left coset and right coset of the subgroup H4 are equal whence
H4 = {ρ0 , ρ1 , ρ2 } is a normal subgroup of D3 .
Group Theory 35
TABLE 2.8:
Factor group of D3
G
H4 H4 H4 µ3
H4 H4 H4 µ3
H4 µ3 H4 µ3 H4

TABLE 2.9:
C2 and Z2
composition
table
e a
e e a
a a e

2.4 Factor Group


G
For a normal subgroup the quotient set N is termed the factor group. The
normal subgroup plays the role of identity which is the unit element in the
factor group. {ρ0 } is a trivial subgroup of D3 and hence is obviously normal.
The factor group of D3 by {ρ0 } is D3 itself. Thus, the factor group D D3 is of
3

order 1 or {ρ0 }.
For D3 we saw {ρ0 , µ1 }, {ρ0 , µ2 }, {ρ0 , µ1 } hence order two subgroups are
not normal subgroups, but {ρ0 , ρ1 , ρ2 } is a normal subgroup. Hence D3 =
{(ρ0 , ρ1 , ρ2 ), (µ1 , µ2 , µ3 )} = H4 ∪ H4 µ3 (or H4 µ2 , H4 µ1 ).
For the normal subgroup H4 the composition table of the factor group is
given in Table 2.8.
Note that it is identical and isomorphic to the composition table of the
cyclic group C2 or (Z2 ) which is given in Table 2.9.
Here (H4 µ3 ) = H4 as µ23 = ρ0 , ρi µ3 ρj µ3 = µ23 ρk ρl = ρk ρh ∈ H4 .
Note the remarkable parallelism between the factor group of D3 as given
in Table 2.8 and Table 2.4 in terms of the rotation ρ1 and the mirror reflection
µi .

————————————————–
Problem 2.17: Study the normal group structure of the {e, a} subgroup
of the Viergruppe V = {e, a, b, c}.
————————————————–
Problem 2.18: Determine the order of the factor group Z6 / < 3 > where
< 3 >= {0, 3}.
————————————————–
36 Group Theory in Particle, Nuclear, and Hadron Physics

Problem 2.19: For the dihedral group D4 given three subgroups H1 =


{ρ0 , ρ2 }, H2 = {ρ0 , µ1 } and H3 = {ρ0 , ρ1 , ρ2 , ρ3 },which are normal subgroups
of D4 ?
————————————————–

For a normal subgroup (also called an invariant subgroup), Hs = sH for


s ∈ G. This implies s−1 Hs = H. This forms a group. A group may be parti-
tioned into classes just as a group may be partitioned into left or right quotient
sets. Consider a, b ∈ G. If there exists an s ∈ G such that

a = s−1 bs

a is said to be conjugate to b by the element s, b = sas−1 and b is conjugate


to a by s−1 . If b and c are conjugate to a, then b and c are conjugate to each
other.
a = s−1 bs, a = t−1 ct ; t, s ∈ G. Then

b = sas−1 = s(t−1 ct)s−1 = (st−1 )c(ts−1 ) = (ts−1 )c(ts−1 ) = x−1 cx

And so b is conjugate to c.
All the elements conjugate to each other form a class. A class may be
denoted by a single element.
Every member of a group G is a member of some class a ∈ G, whence
e−1 ae ∈ G and so it is in a class constructed from itself. The identity e of G
always forms a class of its own. As,

x ∈ G x−1 ex = x−1 x = e

Also note that no element of G can be a member of two different classes.


So if a ∈ G is a member of a class containing b and is a member of a class
containing c, then a is a conjugate of b and a is a conjugate of c, hence b and
c are conjugate to each other. And so these belong to the same class.
Consequently every group can be broken up into separate disjoint classes.
Order of elements in a class in a divisor of o G, the order of G. Thus if o G = 6,
then it can have classes of 1,2,3 or 6, order.
Now let us study classes of the D3 group. It has element {ρ0 , ρ1 , ρ2 , µ1 ,
µ2 , µ3 }. ρ0 as an identity and a class by itself. Next class of [µ3 ] is

ρ−1
0 µ3 ρ0 = µ3 , µ−1
3 µ3 µ3 = µ3

µ−1
1 µ3 µ1 = µ2 , ρ−1
1 µ3 ρ1 = µ2

µ−1
2 µ3 µ2 = µ1 , ρ−1
2 µ3 ρ2 = µ1
Group Theory 37
TABLE 2.10:
Z3 composition
table
0 1 2
0 0 1 2
1 1 2 0
2 2 0 1

Whence the class of [µ3 ] is [µ3 ] = [µ1 , µ2 , µ3 ]

Similarly had we taken µ1 or µ2 ; [µ1 ] and [µ2 ] would have given the same
class structure. Thus, a class can be identified by any one of its member-
elements. Next the class of [ρ1 ] is

ρ−1
0 ρ1 ρ0 = ρ1 , ρ−1
1 ρ1 ρ1 = ρ1

µ−1
1 ρ1 µ1 = ρ2 , ρ−1
2 ρ1 ρ2 = ρ1

µ−1
2 ρ1 µ2 = ρ2 , µ−1
3 ρ1 µ3 = ρ2

And the class of ρ1 is [ρ1 ] = [ρ1 , ρ2 ]

Thus, D3 has three disjoint classes [ρ0 ], [ρ1 , ρ2 ], [µ1 , µ2 , µ3 ] of ordered 1,2,3
which are divisors of 6.

————————————————–
Problem 2.20: Prove that for an Abelian group every element is a class
by itself.
————————————————–

2.5 Homomorphism and Isomorphism


The composition table of a finite group has all the information of its group
structure. If we have two composition tables of two different groups of the
same order, then different “names” of its members like {1, ω, ω 2 } for cube
roots of unity occur, while for C3 the names of the group, again of order 3, are
{ρ0 , ρ1 , ρ2 }. If we exchange the pair of names 1 ↔ ρ, ω ↔ ρ1 , ω 2 ↔ ρ2 then
the two composition tables are identical. Also for Z3 the composition Table
2.10 is one only.
Nexy we now exchange these names if the groups [Z3 , cuberoots, C3 ]
38 Group Theory in Particle, Nuclear, and Hadron Physics
TABLE 2.11:
Order three group
e a b
e e a b
a a b e
b b e a

{(0 ↔ 1 ↔ ρ0 ), (1 ↔ ω ↔ ρ1 )} then all these can be referred to the same


abstract group with arbitrary names (e, a, b) as shown in the Table 2.11.
So as far as the group structure goes, all three different entities Z3 , cube
roots, and C3 are identical or are exactly the same. Thus we state that all
three groups are isomorphic to each other.
As another example, the composition table of the 4th roots of unity has
elements {1, i, −1, −i}. Compare it to the Z4 (addition modulo4) composition
table. If we exchange the names of the pair {Z4 , 4th roots} as {0 ↔ 1, 1 ↔
i, 2 ↔ −1, 3 ↔ −i} then one notices that these are exactly the same and
hence these groups of order 4 are isomorphic. How about the other order 4
group, the Viergruppe V ? Clearly it is not isomorphic to Z4 and is thus an
independent and a different group of order 4.
Let us generalize, given two groups of the same order of G =
{e, a, b, c...}; G0 = {e0 , a0 , b0 , c0 ....}. These two groups G and G0 are isomor-
phic if there exists a one-to-one correspondence between the elements of G
and G0 . This is that all the elements of the two groups are related by the
simultaneous exchange of all the elements {a ↔ a0 , b ↔ b0 , c ↔ c0 , ....}. Thus,
a multiplication of ab = c in G means a0 b0 = c0 in G0 must hold. Also note
that the identities e ↔ e0 map one-to-one too. Thus, these have identical (or
the same) composition tables. This is isomorphism.
If the mapping between G and G0 is many-to-one, then the correspondence
between the two is referred to as homomorphism. There is a homomorphism
from a group G to G0 if to each element a in G there corresponds a unique
element φ(a) of G0 such that φ(ab) = φ(a)φ(b). This mapping φ must be defined
for each element of G. Thus, φ(a) in G0 is the map of elements of G under
the homomorphism. Clearly several elements of G may be mapped onto the
same element of G0 . So even though a 6= b in G, φ(a) = φ(b) in G0 . Thus, one
may have n-to-one mapping in the homomorphism from G to G0 . Clearly for
n = 1 the homomorphism reduces to an isomorphism.
Thus, we take 4th roots of unity G0 = {1, i, −1, −i} and the second group
0
G = {1, −1} as the square roots of unity under multiplication. We map these
as
φ
G → G0

{1, −1} → 1 = φ(1) = φ(−1)


Group Theory 39

{i, −i} → −1 = φ(i) = φ(−i)

This mapping is a 2-to-1 homomorphism. Note it should satisfy,

φ(g1 )φ(g2 ) = φ(g1 g2 )

This is satisfied as

φ(1)φ(−1) = (1)(1) = 1 = φ(−1) = φ((0)(−1))

φ(−1)φ(i) = (1)(−1) = −1 = φ(−i) = φ((−1)(i))

and so on.
Next let us take G = {(e1 , e2 , ....en ), (a1 , a2 , ...an ), ...}. Here say e1 is the
identity of G. This group is of order ng where another group G0 = {e1 a1 b1 c1 ...}
is of order g. If we can split G into g sets (ei ), (ai ) where each contains n ele-
ments, such that the elements of G can be mapped onto the element of G as

e1 e2 , ....en → e ; a1 , a2 , ...an → a, etc.

then the group G is said to be homomorphic to G0 if the mapping is such that


the product

ai bj = ck , 1 ≤ k ≤ n

In G this implies ab = c in G0 when c is the map on G0 of the element


(c1 , c2 , c3 ....) of G (and vice versa). This is thus an n-to-1 homomorphism of
G onto G0 .
If this is so, then note the {ei } set of G maps onto the identity e of G0 .
For homomorphism to hold for the whole G → G0 structure, it is essential
that {ei } be a normal subgroup of G. {ei } is denominated the kernel of the
homomorphic mapping φ : G → G0 . In the example above, the kernel is the set
{1, −1}. The kernel is not only a subgroup of G but it is a normal subgroup
of G.
As we demonstrated above, given a normal subgroup N of a group G, N ⊂
G
G, we can form a factor group N consisting of elements N, aN, bN, · · · . Here
N plays the role of the “identity” element of the factor group. Thus, taking
N as the kernel of the mapping, the factor group is thus the homomorphism
G
φ:G→ N . Now as only normal subgroups serve as a kernel of homomorphic
mappings and as each normal subgroup can act as a kernel one may generalize
by stating that each and every homomorphic mapping of G is isomorphic to
some factor group of G0 . This is termed the Isomorphism Theorem. This
is one of several isomorphism theorems in literature. Hence, the existence
of the normal subgroups and the corresponding factor groups is sufficient to
determine all the homomorphic mappings.
40 Group Theory in Particle, Nuclear, and Hadron Physics

2.6 Torsion Group and Betti Number


We have come across several Abelian groups; for example, Z under the
binary operation multiplication, C complex number under addition, Z3 and
cube roots of unity, Z4 and so forth. We can consider these as being alike in
as much as all of these are Abelian. But how can these be classified so as to
be able to distinguish between them?
An element a of a group G with identity e has an order r > 0 if ar = 0 and
no smaller positive power of a is the identity. Take Q, the Abelian additive
group of rationals. Note that every element of Q except 0 is of infinite order.
Next in Q m
Z , every element is of finite order. If q ∈ Q, q = n where m, n ∈ Z,
Q
so n(q + Z) = nq + Z = m + Z = Z and so Z is of finite order for every
element. The complex number C can have both finite order and infinite order
elements. Identity of C is 1. As (−1)2 = 1 it implies that (−1) is of order two,
thus it is of finite order. In C if 3q = 1 is only true for q = 0 and thus 3 is of
infinite order.
If we have a group G in which every element is of finite order, then it is
called a Torsion group, like the example of Q Z above. If in a group G every
element other than the identity is of infinite order, then G is said to be Torsion-
free (e.g., Q above). If both hold, and an element of infinite order and an
element (not equal to identity) of finite order exist in a group G, then it is
termed a mixed group (e.g., C above).
We prove that there are two groups of order four – one, the 4th roots of
unity which is isomorphic to Z4 , and the other one – the Viergruppe V . Can
we have V as isomorphic to some cyclic groups – if not to a single group? The
product of a cyclic group allows us to obtain new groups – in fact, a larger
class of Abelian groups which includes all the Abelian groups of finite order.
External Direct Product is a constructive method of utilizing already
known finite groups to construct more new groups. First one defines a Carte-
sian product of sets, S1 , S2 ....Sn as the set of ordered n-tuplets (a1 , a2 , ....an )
where a :∈ Si . It is written as

S1 × S2 × · · · Sn = Xni=1 Si .

Now given a group G1 , G2 .....Gn ; we can form the product group


Xni=1 Gi . Let us define binary operations of multiplication by components.
For (a1 , a2 .....an ) and (b1 , b2 .....bn ) in Xni=1 Gi define

(a1 , a2 ......an )(b1 , b2 .....bn ) = (a1 b1 , a2 b2 , .......an bn ).

Thus, Xni=1 Gi is an external direct product group of Gi under the above


binary operation.
Group Theory 41

1. Closure: ai ∈ Gi , bi ∈ Gi , with Gi a group; ai bi ∈ Gi and hence the


above operation is closed on Xni=1 Gi ; satisfied.

2. Associativity: (a1 , a2 · · · an ) [(b1 , b2 , · · · bn )(c1 , c2 , · · · cn )]

= (a1 , a2 , · · · an )(b1 c1 , b2 c2 , · · · bn cn )

= a1 (b1 c1 ), a2 (b2 c2 ), · · · an (bn cn )

= (a1 b1 )c1 , (a2 b2 )c2 · · · (an bn )cn

= [(a1 , a2 , · · · an )(b1 , b2 , · · · bn )](c1 , c2 , · · · cn ); satisfied.

3. Identity: If ei is the identity of Gi , then multiplication by components


(e1 , e2 , · · · en ) is an identity in Xni=1 Gi ; satisfied.

4. Inverse: (a1 , a2 , .....an )(a−1 −1 −1


1 , a2 ......an )

= (a1 a−1 −1 −1
1 , a2 a2 , .......an an )

= (e1 , e2 ........en ) satisfied.

And the latter is the inverse of the former in the above product.

n
Thus, Xi=1 Gi is a group
If the operation on each G Pi nis commutative, then we have the external
direct sum of Gi expressed as i=1 Gi . Clearly the external direct product of
Abelian groups is itself an Abelian group.
Examples:
(a) Consider Z2 × Z2 ,. Hence it has 2.2 = 4 elements as Z2 = {0, 1}. So
Z2 × Z2 has elements {(0, 0), (0, 1), (1, 0), (1, 1)}.
As the operation is additive for Z2 and so too for the external direct prod-
uct Z2 × Z2 . Is Z2 × Z2 cyclic? We have only to find a generator. Let us try
the generator (1, 1):

(1, 1) = (1, 1) ; 2(1, 1) = (1, 1) + (1, 1) = (0, 0)

Hence, the generator gives back identity only after two operations, while
in a 4-order cyclic group, a generator would yield the identity only after 4
operations. And hence Z2 × Z2 is not cyclic. As to the group of order 4, we
already know that Z4 = C4 is a cyclic group. And hence Z2 × Z2 should be
isomorphic to the Viergruppe V . Remember that there are only two indepen-
dent groups of order 4. So all the groups of order 4 should be identical to or
be isomorphic to either Z4 or V .
(b) Take the group Z2 × Z3 with 2.3 = 6 elements and which are
42 Group Theory in Particle, Nuclear, and Hadron Physics

{(0, 0), (0, 1), (0, 2), (1, 0), (1, 1), (1, 2)}. If it is cyclic there would be a gen-
erator that generates it. Consider (1,1) as that element. Then,

(1, 1) = (1, 1) ; 2(1, 1) = (1, 1) + (1, 1) = (0, 2)

3(1, 1) = (0, 2) + (1, 1) = (1, 0) ; 4(1, 1) = (1, 0) + (1, 1) = (0, 1)

5(1, 1) = (0, 1) + (1, 1) = (1, 2) ; 6(1, 1) = (1, 2) + (1, 1) = (0, 0)

Consequently it is a cyclic group of order 6. Hence (1,1) generate the hole


of Z2 × Z3 . Since there is only one such cyclic group of a given order up to
isomorphism, thus (Z2 × Z3 ) = Z6 and so the two are isomorphic.

————————————————–
Problem 2.21: Is Z3 × Z3 isomorphic to Z9 ?
————————————————–

Thus Zi × Zj is not always isomorphic to the group Zij . So when is it so?


As per the following theorem

Theorem : If i and j are relatively prime; i.e., if the GCD of i and j is 1,


then Zi × Zj is isomorphic to Zij .

To prove this consider the cyclic subgroup of Zi ×Zj generated by (1, 1). We
know that the order of this cyclic subgroup is the smallest power of (1,1) which
yields the identity (0, 0) of Zi × Zj . So let us add (1, 1), respectively. 1 ∈ Zi
yields 0 after i summations, 2i summations, and so forth. And 1 ∈ Zj gives
0 after j, 2j, · · · summations. So to obtain (0,0) the number of summations
must be the multiple (ij). Thus, the GCD of i and j should be 1. Then (1,1)
generates a cyclic subgroup of order (ij), which is the order of the whole group
Zi × Zj .
The above can be generalized. If number im for m = 1, .....n are such
that a GCD of any two of these is number 1, then Xnm=1 Xim is cyclic and is
isomorphic to Zi1 ,i2 ,.....in
If i = (p1 )i1 (p2 )i2 ......(pr )in where pi are distinct prime numbers, then Zi
is isomorphic to

Z(p1 )i1 × Z(p2 )i2 × ...........Z(pr )ir

Given Z72 as (8, 9) =(23 , 32 ) are relative prime, so it is isomorphic to


Z8 × Z9 .
Given a ∈ G for a group. If an = e for some positive integer n the least
positive integer, as we saw earlier, is called the order of a. If it turns out that
there is no such n, then a is of infinite order. Clearly a is an element of a group
Group Theory 43

G, the order of a is the same as the order of the cyclic subgroup generated by
a.
We can now generalize the product Zi × Zj and Zij as per the theorem
above to Zi × Zj × ......Zr and Zij........r , then clearly the product is the least
common multiple of (i, j, .....r).
Next if Xni=1 Gi is an external direct product of groups Gi , then there ex-
ists a subset

G∗i = {(e1 , e2 , ......, ei−1 , ai , ei+1 ...........em ) | ai ∈ Gi }

This is the set of all n-tuples with the identity element in all locations ex-
cept the ith which gives the element ai , and must be a subgroup of Xni=1 Gi. .
If we define a mapping,

πi (e1 , e2 , .......ei−1 , ai , ei+1 ..........em ) = ai

Then this yields the isomorphic mapping to Gi . One considers Xni=1 Gi


as the internal direct product of these subgroups G∗i . Here the word internal
implies that we are regarding the component groups as subgroups of the prod-
uct group. However usually it will be clear from the context of the discussion
whether we are talking about the internal or the external direct products.

————————————————–
Problem 2.22: Derive all the elements of the group Z2 × Z4 . Is the group
cyclic in nature?
————————————————–
Problem 2.23: Of all the orders of the cyclic subgroups of (a) Z6 × Z8
and (b) Z12 × Z15 , which is the largest order?
————————————————–

Hence in our above examples Q Z is a torsion group and Q is a torsion-


free group. Hence every finite group is a torsion group. Z under addition is
torsion-free.
G
Note that if G is a group, H a subgroup of G such that H is torsion-free,
then one can see that H contains the torsion subgroup of G. Assume g ∈ G
G G
is of finite order. Then g + H is also finite order in H . We know that H is
torsion-free whence g + H = H. Thus, we have g ∈ H. Hence every element
of finite order in G is also contained in H. That is, the torsion subgroup of G
exists in H.
The group Z × Z2 is generated by {(1, 0), (0, 1)}. The element (0,1) is of
order 2 and the element (1,0) is not of finite order. So what is the torsion
subgroup of Z × Z2 ? It is T = {(0, 0), (0, 1)}. Note that the full group Z × Z2
is mixed and only its subgroup T is an Abelian torsion subgroup.
44 Group Theory in Particle, Nuclear, and Hadron Physics

————————————————–
Problem 2.24: Let a group K be a direct sum of torsion groups. Prove
that K is a torsion group too.
————————————————–
Problem 2.25: Given R, the group of real numbers under addition and
Z, a group of integers, determine the torsion subgroup of R
Z.
————————————————–
Problem 2.26: Is Z25 × Z9 isomorphic to Z5 × Z5 × Z9 ?
————————————————–

Given a collection of sets {ST i | i ∈ I}, the intersection of the sets Si


is in standard notation given as i∈I Si . Clearly the intersection of a set of
subgroups Hi , i ∈ I of a group G is again a subgroup of G. Take an element
ai G for i ∈ I where I lists the element ai . Thus there is at least one subgroup
of G which contains all ai , and that is the group G itself. Now this ai should
be part of the smallest subgroup of intersection of all the subgroups of G
containing ai . Thus, this smallest subgroup is generated by {ai | i ∈ I}. If G
is all of the above subgroup, then {ai | i ∈ I} generates G and we say that
the ai are generators of G. It may be that G is finitely generated if a finite set
{ai | i ∈ I} exists.
For Abelian group G if ai ∈ G for i ∈ I, then the subgroup H ⊂ G
generated by {ai | i ∈ I} has elements of G which are finite products of
integral powers of ai .
For example, there may be elements (a1 )3 (a2 )7 . One may understand this
by assuming that there is a set K ⊆ H which consists of all finite products of
integral powers of ai . Clearly any product in K is in K itself. Also (a1 )0 = e.
Given k in K, k −1 exists for example if (a1 )3 (a2 )7 its inverse is (a2 )−7 (a1 )−3
and is in K. But as H is the smallest subgroup, K = H.
We define a torsion group T as one in which every element is of finite
order. The group F is torsion-free only if the identity is of finite order and all
other elements are of infinite order.
Thus given a ∈ T and an = e then
e = en = (aa−1 )n = an (a−1 )n = e(a−1 )n = (a−1 )n
whence a−1 is of finite order and thus is in T .
We can define mixed groups as follows. If G is a finitely generated Abelian
group with a torsion subgroup T , then there must exist a torsion-free subgroup
F of G such that G is a direct product of T × F .
Now this breakup of any Abelian group into a torsion part and a torsion-
free component leads to defining two fundamental quantities in algebraic
topology (Appendix G) – the Betti number and the torsion coefficients [22],
[21], [23], [24].
If a finitely generated torsion-free Abelian group F is isomorphic to Z ×
Z × · · · × Z for m number of factors, then m is a fundamental and unique
number defining F . It is called the Betti number of F .
Group Theory 45

The Abelian torsion group component may exhibit isomorphism with re-
spect to two different types of direct products of cyclic groups,

(A) Z(p1 ) x1 × Z(p2 ) x2 × ......... × Z(pn ) xn

Where pi (even repeating) primes.

(B) Zn1 × Zn2 × ......... × Zns

where ni divides ni+1 . The number ni uniquely defines the torsion coefficients
of T .
The above provides a complete description, up to an isomorphism of all
finite Abelian groups. Note that the T (defined by unique torsion coefficients
as above) and F (defined by the Betti number as above), therefore provide
information of all that can be known about the finite Abelian group.
Let us now discuss a practical way of determining the Betti number and
the torsion coefficients of a given direct product of groups. We write different
primes starting at number two in a product group in different lines in ascend-
ing magnitude, one line below the other, while ensuring alignment with the
last column. Then the product of numbers in each column yields the corre-
sponding isomorphic group. So given Z2 × Z4 × Z3 × Z3 × Z5 we formed the
following lines

2 4
3 3
5
6 60
Thus, the above product group is isomorphic to the product group Z6 ×Z60
So its torsion coefficients are 6 and 60.
Let us find all Abelian groups of order 360. As the product of primes
360 = 23 .32 .5. Then on the basis of the above discussion, the following six
combinations (up to isomorphism) are possible. We obtain the torsion coeffi-
cients for each below.

(1) Z2 × Z2 × Z2 × Z3 × Z3 × Z5

As per the above rule we write the coefficient in lines as follows:

2 2 2
3 3
5
2 6 30
Thus the above group is isomorphic to Z2 × Z6 × Z30 and its torsion coef-
46 Group Theory in Particle, Nuclear, and Hadron Physics

ficients are 2,6,30.

(2) Z2 × Z4 × Z3 × Z3 × Z5 discussed above as Z6 × Z60

(3) Z2 × Z2 × Z2 × Z9 × Z5
2 2 2
5
9
2 2 90
→ Z2 × Z2 × Z90

(4) Z2 × Z4 × Z4 × Z5
2 4
5
9
2 180
→ Z2 × Z180

(5) Z8 × Z3 × Z3 × Z5
3 3
5
8
3 120
→ Z3 × Z120

(6) Z8 × Z9 × Z5 → Z360

————————————————–
Problem 2.27: Find all Abelian groups (up to isomorphism) of order
1089. Obtain the torsion coefficients.
————————————————–
Problem 2.28: Derive the torsion coefficients and Betti number of the
product group,

Z6 × Z × Z10 × Z × Z12 × Z

————————————————–

One more look at the finite group, the additive group of rationals modulo
the integers. We already have seen that it is a torsion group.
Group Theory 47

(Q
Z ) = {x + Z | x + Z of order a power of p }

= {x + Z | ps x ∈ Z}

{ pns + Z | for integers and 0 ≤ n < ps−1 }

Now using addition notation below (instead of the product in defining the
torsion coefficients),
Q P Q
Z = p∈π ( Z )p

where π is the set of all primes.


Let us just now use the ideas of left cosets and round subgroups for the
direct product of groups. Let us compute the number of left cosets of the
subgroup, < 0 >< 1 >< 2 > of Z3 × Z2 × Z4 .

Z3 × Z2 × Z4 or {0, 1, 2} × {0, 1} × {0, 1, 2, 3}

So there are 24 terms like

{(0, 0, 0), (0, 0, 1), (0, 0, 2), (0, 0, 3), (0, 1, 0), (0, 1, 1), (0, 1, 2), (0, 1, 3) · · · }

Now

< 0 >< 1 >< 2 > or {0}×{0, 1}×{0, 2} ≡ (0, 0, 0), (0, 0, 2), (0, 1, 0), (0, 1, 2)

There are these four elements in the subgroup and so the left coset of the
above group is 24
4 = 6.

————————————————–
Problem 2.29: Determine the number of left cosets of the subgroup,
< 1 >< 0 >< 0 > of Z3 × Z2 × Z4 .
————————————————–
Problem 2.30: Find the left cosets of the subgroup < 18 > of Z36 .
————————————————–
Problem 2.31: Given the cyclic subgroup < (1, 2) > of Z2 × Z4 , find its
left cosets. Does it form a group? If yes, what is it isomorphic to?
————————————————–
48 Group Theory in Particle, Nuclear, and Hadron Physics

2.7 Appendix B: Matrices


Display a rectangular array of expression Aij with i varying from integers
l to m and j varying from l to n as:
 
A11 A12 . . . A1n
 A21 A22 . . . A2n 
 
A=  . . (2.9)


 . . 
Am1 Am2 . . . Amn
with elements Aij occupying the position of intersection of the ith row and
the j th column. A is called a matrix. Matrices A and B add or subtract only
if they have the same number of rows and columns.

A±B =C (2.10)
with elements Aij ± Bij = Cij
A scalar α multiplying a matrix A yields

αA = B (2.11)
with each element of A multiplied by the number (real or complex)

αAij = Bij (2.12)


Multiplication of matrices A and B is displayed as:

AB = C (2.13)
Pn
Ai1 A1j + Ai2 A2j + . . . + Ain Bnj = k=1 Aik Bkj = Cij
n
X
(AB)ij = Aik Bkj (2.14)
k=1

Written explicitly the elements Cij are the sum of the products of corre-
sponding terms in the ith row of the first matrix and the j th column of the
second row

  
. . . . . B1j .  
 Ai1 . . .
Ai2 . Ain   . B2j .  

AB = 
 .
 = . Cij .  (2.15)
. . .  . . . 
. . .
. . . . . Bnj .

Note that it is essential that for the matrix product to be defined properly,
the number of columns in A must equal the number of rows in B. Note that
Group Theory 49

the matrix product in general need not be commutative, i.e., AB need not
equal BA. If as a special situation AB = BA, then the matrices are said to
commute [25], [23], [26].
The matrix product obeys the associative law: (AB)C = A(BC), i.e., it
does not matter which two of the three matrices ABC are multiplied first.
Only the order of the matrices ABC is important.

————————————————–
Problem B.1: Prove that (AB)C = A(BC) where bracket indicates which
product is taken first.
————————————————–

We define the unit (or identity) matrix I as a square matrix with ele-
ments δjk (the Kroenecker delta function), with
X
(IA)ij = δik Akj = Aij
k

and
X
(AI)ij = Aik δkj = Aij (2.16)
k

Hence it is called a unit matrix, that is, it does not matter whether we
multiply with it from the left or from the right.
Diagonal matrix is a matrix with elements Aij δij . So
 
A11 O . . O
 O A22 . . O 
Ann = [Aij δij ]nn =  .
 (2.17)
. . . . 
O . . . Ann

The matrix product of a diagonal matrix A with another matrix D of the


same order yields a matrix with elements:
X
AD = Aik δik Dkj = Aii Dij (2.18)
k

Determinant of a matrix
Every square matrix has a real number associated with it which is called
its determinant. First define a cofactor matrix as

Ac = [Aij ] with Acij = (−1)i+j Mij (2.19)


where Mij is the minor of the matrix which is obtained by forming a deter-
minant with ith row and j th column of A omitted. So for
50 Group Theory in Particle, Nuclear, and Hadron Physics

 
a11 a12 a13
A =  a21 a22 a23  (2.20)
a31 a32 a33


a a13 a a23 a a12
M31 = 12 , M11 = 22 , M23 = 11 , (2.21)
a22 a23 a32 a33 a31 a32

Thus the determinant is defined as:


N
X N
X
det(A) = |A| = Aij Acij = (−1)i+j Aij Mij (2.22)
j=1

We define the inverse matrix A−1 as per the conditions

A−1 A = AA−1 = I (2.23)


where I is the n × n identity matrix. An inverse matrix exists only if det(A)
is non-zero. Then,
Acji
(A−1 )ij = (2.24)
det(A)
where Acji is the transpose of the cofactor matrix. Then
X X Acjl
(AA−1 )ij = Ail A−1
lj = Ail = δij (2.25)
det(A)
l l

We can show this for A−1 A in the same manner.

————————————————–  
1 3
Problem B.2: Find the inverse of the matrix A=
2 1
————————————————–

We define the following special matrices of a matrix A with elements Aij


∼ ∼
Transpose: A, (A)ij = Aij
Complex conjugate: A∗ (A∗ )ij = (Aij )∗
∼ ∗
Adjoint or Hermitian: conjugate A† = (A) so (A† )ij = (4ji )∗

The following conditions specify some special mathematical definitions:

Diagonal: if Aij = 0 f or i 6= j
Idempotent: if A2 = A

Symmetric: A=A
Group Theory 51

Skew-symmetric (or anti-symmetric): A = −A
Real: A∗ = A∼
Orthogonal: A = A−1
Hermitian (or self-adjoint): A† = A
+
Anti-Hermitian: A = −A
Unitary: A+ = A−1
Anti-unitary: A+ = −A−1

Note that if A is m × n matrix, then A and A† are n × m matrices and

A is m × n. Also if a matrix is both orthogonal and real, it is automatically
unitary and a matrix that is symmetric and real is automatically Hermitian.
Some important properties of determinants:


det A = det A
det A∗ = (det A)∗ (2.26)
det (AB) = (detA) (detB)
∼ ∼
Now for the orthogonal matrix A = A−1 and as A−1 A = I so AA = I.
Then,

∼ ∼
det (AA) = (det A)(det A) = (det A)(det A) = (detA)2 = I (2.27)

Hence det A = +1 or − 1
For the unitary matrix A−1 = A† , so A−1 A = A† A = I. And so


det (A† A) = (detA† )(detA) = det(A)∗ det A

= (detA)∗ det A = (detA)∗ (detA) (2.28)

= |det A|2 = 1

So det A = eiα where α is a real number


Trace of a square matrix is the sum of diagonal elements.
X
Tr A = Aii (2.29)
i

If T r A = 0, then the matrix is called traceless.


Note Tr(A+B)=TrA+TrB

————————————————–
Problem B.3:
(a) Show that T r(AB) = T r(BA)
52 Group Theory in Particle, Nuclear, and Hadron Physics
∼ ∼
(b) ^ = B.A
(A.B)
(c) (A.B)∗ = A∗ B ∗
(d) (A.B)† = B + A†
(e) If A and B are non-singular (det(A) and det(B) are both nonzero)
then (AB)−1 = B −1 A−1
(f) If A and B are Hermitian matrices then i(AB-BA) is also Hermitian.
————————————————–

Similarity transformation: If A and A0 are n × n matrices then these


are said to be related by a similarity transformation:

A0 = S −1 AS (2.30)
where S is a n × n non-singular matrix.

————————————————–
Problem B.4: For the above similarity transformation show that T rA0 =
T rA.
————————————————–

Direct Product or Kroenecker Product

C(mp)×(nq) of two matrices Am×n and Bp×q of different orders is defined


as:

C(mp)×(nq) = Amn ⊗ Bpq (2.31)


The direct product in components is written as:

Cik ,jl = Aij Bkl (2.32)


Example:
 Given 2× 2 matrices
 
A11 A12 B11 B12
A= B=
A21 A22 B21 B22
Then the direct product
 
A11 B11 A11 B12 | A12 B11 A12 B12
   A11 B21 A11 B22 | A12 B21 A12 B22 
A11 B A12 B  
A⊗B =
A21 B A22 B
=  − − − − −  
 A21 B11 A21 B12 | A22 B11 A22 B12 
A21 B21 A21 B22 | A22 B21 A22 B22

————————————————–
Problem B.5: Dirac derived matrices for this 4-dimensional equation as:
σi ⊗ σj ; i, j = 0, 1, 2, 3 ( σi are 2 × 2 Pauli matrices, i=1,2,3 and σ0 is 2 × 2
identity matrix ). Find σ1 ⊗ σ3 .
————————————————–
Group Theory 53

The direct product of matrices is associative but not commutative. It has


the following properties:

A ⊗ (B + C) = A ⊗ B + B ⊗ C (2.33a)
(A ⊗ B)(C ⊗ D) = AC ⊗ BD (2.33b)

————————————————–
Problem B.6: Show that if A(n×n) and B(m×m) are both unitary, then
A ⊗ B is also unitary.
————————————————–

Note that a square matrix can be represented as a sum of a Hermitian and


an anti-Hermitian matrix of the same order:

1 1
A= (A + A† ) + (A − A† )
2 2 (2.34)
=H +K

Where H is Hermitian H † = 21 (A + A† )† = 21 (A† + A) = H

and K is anti-Hermitian K † = 12 (A − A† )† = 21 (A† − A) = −K


Note that matrix inversion gives a simple method for solving a system of
non-homogeneous equations. For example:

x + 3y = 2
2x + 3y = 3

Written in matrix from AX = C with


     
1 3 x 2
A= X= ,C=
2 1 y 3
 1 3

−1 −5 5
From Problem B.2 we see that A = 2
5 − 15
 1 3
    7

−5 2
so X = A−1 C = 2
5 = 5
5 − 15 3 1
5

hence x = 57 , y= 2
5
54 Group Theory in Particle, Nuclear, and Hadron Physics

2.8 Solutions of Problems


Solution 2.1:
For a, b, c, e ∈ G with e an identity. Assume ab = e and ac = e with b 6= c;
i.e., two different elements give an identity. However, then ab = ac, a−1 ab =
a−1 ac leading to b = c. But this is a contradiction as all the elements of the
set in G are defined to be different. So the identity is unique.
————————————————–
Solution 2.2:
(1) Closure: a, b ∈ Z ; a, b ∈ Z; satisfied.

(2) Associativity: (a· b)· c = a· (b· c) satisfied.

(3) Identity: 1· a = a· 1; so 1 is multiplicative identity; satisfied.

(4) Inverse: a· a−1 = 1; a−1 = a1 ; for a ∈ Z; a1 ∈


/ Z; not satisfied so Z is not
a group under multiplication.
————————————————–
Solution 2.3:
a1 a1 a1 a2
(1) Closure: b1 . b1 = b1 b2= ab ∈ Q; satisfied.
   
a1 a2 a3 a1 a2 a3
(2) Associativity: b . b2 . c3 = b1 . b2 . b3 ; satisfied.

(3) Identity: 1 is identity 1. ab = ab . 1; satisfied.

a b a b
(4) Inverse: b ∈ Q0 , a is inverse as b · a = 1; satisfied.

So it is a group under multiplication. Now with 0 included, the set of all


rational number Q steps 1,2,3 hold but the inverse of 0 does not exist as
a
0 ∈
/ Q. So it is not a group anymore.

————————————————–
Solution 2.4:
(1) Closure: α + β = (a1 + ib1 ) + (a2 + ib2 ) = (a1 + a2 ) + i(b1 + b2 ) ∈ C;
satisfied.

(2) Associativity: (α + β) + γ = α + (β + γ) ∈ C; satisfied.

(3) Identity: 0 + i0 is identity; satisfied.


Group Theory 55
TABLE 2.12:
Composition
table of square
roots of unity
1 -1
1 1 -1
-1 -1 1

TABLE 2.13:
Composition table of
fourth roots of unity
1 i -1 -i
1 1 i -1 -i
i i -1 -i 1
-1 -1 -i 1 i
-i -i 1 i -1

(4) Inverse: −a − ib is inverse of a + ib ∈ C; satisfied.

Hence it is an infinite Abelian group.


————————————————–
Solution 2.5:
Square roots are (1,-1) and the composition table is in Table 2.12.
i
The 4th roots of unity are: e2πr 4 −→ (1, i, −1, −i), +1 is the identity and
its inverses are (1, −i, −1, +i), respectively. Its composition table is in Table
2.13.
————————————————–
Solution 2.6:
ab
(1) Closure: a • b = 2 ∈ Q+ ; satisfied.

(2) Associativity: (a • b) • c = ( ab
2 )•c=
abc
4 = a • (b • c); satisfied.

2a
(3) Identity: 2 • a = a • 2 = 2 = a ∈ Q+ , so 2 is identity; satisfied.
4 4
a•( a ) (a )•a
(4) Inverse: 2 = 2 = 2; So a−1 = a4 for a ∈ Q+ ; satisfied.

So it is a group under the above composition law in Q+ .


————————————————–
Solution 2.7:
       
3 2 2 3
µ3 ρ1 = µ3 = = µ1
1 2 3 1 1 3 1 2
56 Group Theory in Particle, Nuclear, and Hadron Physics
TABLE 2.14: Z6
the addition modulo 6
composition table
0 1 2 3 4 5
0 0 1 2 3 4 5
1 1 2 3 4 5 0
2 2 3 4 5 0 1
3 3 4 5 0 1 2
4 4 5 0 1 2 3
5 5 0 1 2 3 4

       
3 2 1 3
µ1 ρ1 = µ1 = = µ2
1 2 3 1 3 2 1 2
       
3 3 1 3
µ1 µ3 = µ1 = = ρ2
1 2 2 1 2 3 1 2
————————————————–
Solution 2.8: Table 2.14.
————————————————–
Solution 2.9:
Comparing Z4 and the fourth-roots of unity tables, we notice the sameness
between them by replacement (0, 1, 2, 3) ↔ (1, i, −1, −i), respectively. In Z4
the identity element is 0 and its various inverses are

(0)−1 = 0 (1)−1 = 1

(1)−1 = 3 (i)−1 = −i

(2)−1 = 2 (−1)−1 = −1

(3)−1 = 1 (−i)−1 = i

But V is not isomorphic to Z4 and thus V is an independent and a different


group of order 4.
————————————————–
Solution 2.10:
The lattice diagram of D4 is depicted in Figure 2.3.
————————————————–
Solution 2.11:
The group with elements {1, i, −1, −i} can be generated by i (or -i) as
i, i2 = −1, i3 = −i, i4 = 1.
Group Theory 57

D
4

{P0.P.M.M2) {Po,P,P,P3} {P..P.W

{
<P,M> {p,M} {VP2} {
"o'V vv

FIGURE 2.3: Lattice diagram showing the subgroup structure of D4

Also C2 (or square roots of unity) are generated by −1, (−1)2 = 1.


————————————————–
Solution 2.12:
Given 1, 1 + 1 = 2 , 1 + 1 + 1 = 3, ......n × 1 = n −1 + 1 = 0 and so also is
(−1) a generator of Z. So < 1 > and < −1 >→ Z.
————————————————–
Solution 2.13:
Take (ap )n = (an )p = (e)p = e as order of ap is n,

(ap )m = (ap )m 6= ej(o < m < n)

n neither divides p nor m and it does not divide (pm).

pm = nq + r (o < r < n)

and (ap )m = apm = anq+r

= anq ar = ear (anq = (an )q = aq 6= e


= ar 6= e (0 < r < n)

Hence, each element of G can be written as some integral power of ap and


thus ap is a generator of G.
58 Group Theory in Particle, Nuclear, and Hadron Physics
TABLE 2.15: Left
cosets of {0,3} of Z6
0 3 1 4 2 5
0 0 3 1 4 2 5
3 3 0 4 1 5 2
1 1 4 2 5 3 0
4 4 1 5 2 0 3
2 2 5 3 0 4 1
5 5 2 0 3 1 4

Conversely, if ap is a generator of G, let d be the GCD of n and p (then


to show that d = 1).
Now, if possible let d > 1. Then the GCD of n and p is d, (n, p) = d → nd
n n p p p
and dp are positive integers. Now (ap ) d = ap d = an d = (an ) d = e( d ) = e
Therefore, the order of ap < n → ap cannot be a generator of G, which is
against ap being a generator. Whence d = 1.
————————————————–
Solution 2.14:
(p, 6) = 1, so < a > and < a5 > generate the whole group.
a2 = ap , p = 2 GCD of (2,6) is d=2, so nd = 26 = 3
and so elements generated by a2 and {a2 , a4 , a6 = e}.
If a3 = ap , p = 3, d of (3,6) is 3. So nd = 36 = 2
and so {a3 , a6 = e} generates a group of order 2.
————————————————–
Solution 2.15:
Z6 = {0, 1, 2, 3, 4, 5} and it composition table was given in Problem
2.8. Now H = {0, 3}, 1+H = {1, 4}, 2+H = {2, 5}. So Z6 splits up as
(0, 3)|(1, 4)|(2, 5). To indicate this property we can rewrite the composition
table of Z6 as in Table 2.15.
————————————————–
Solution 2.16:
An Abelian group is one in which all the elements commute. Given an
Abelian group

G = {e, g1 , g2 ....gn } the left coset with g ∈ G is

gG = {ge, gg1 , gg2 , · · · ggn }

Right cosets

Gg = {eg, g1 g, g2 g, · · · gn g}

So, gG = Gg and so for all g, G is normal.


Group Theory 59
TABLE 2.16:
Viergruppe V
rearranged
e a b c
e e a b c
a a e c b
b b c e a
c c b a e

TABLE 2.17:
Factor group of
V
V
H H bH
H H bH
bH bH H

————————————————————————————–
Solution 2.17:
Given H ⊂ V with H = {e, a}, note bH = {be, ba} = {b, c} = cH.
Thus V = H ∪ bH. This enables us to rearrange the composition table of
V as in Table 2.16.
Hence H and (cH or bH) form the factor group whose composition table
is given in Table 2.17.
Note that bHbH = H as b2 = e and bhk bhl = b2 hi hj = hi hj ∈ H.
There are actually three normal subgroups of index two here. However,
each has the same factor structure.
V has {e} as a trivial normal subgroup of order four and the factor group
is V itself. Next, the trivial normal subgroup is V itself and the factor group
v
v is of order 1 as {e}.
————————————————–
Solution 2.18:
Refer to Problem 2.15. The order is 26 = 3.
————————————————–
Solution 2.19:
µ1 H3 = {µ1 , δ1 , µ2 , δ2 } = H3 µ1 ; µ2 H3 = {µ2 , δ2 , µ1 , δ1 } = H3 µ2
D4
So D4 = H3 ∪ µ1 H3 given these normal subgroups and so the factor of H 3
is given in Table 2.8.
Now H2 = {ρ0 , µ1 }
So ρ1 H2 = {ρ1 ρ0 , ρ1 µ! } = {ρ1 , δ1 } ; H2 ρ1 = {ρ0 ρ1 , µ1 ρ1 } = {ρ1 , δ2 }
Hence, left and right cosets are not equal and thus do not form normal
subgroups. But H1 = {ρ0 , ρ2 }
60 Group Theory in Particle, Nuclear, and Hadron Physics
TABLE 2.18:
Factor group of D4
H3 µ1 H3
H3 H3 µ1 H3
µ1 H3 µ1 H3 H3

TABLE 2.19: Blocked composition


table of D4
ρ0 ρ2 ρ1 ρ3 µ1 µ2 δ1 δ2
ρ0 ρ0 ρ2 ρ1 ρ3 µ1 µ2 δ1 δ2
ρ2 ρ2 ρ0 ρ3 ρ1 µ2 µ1 δ2 δ1
ρ1 ρ1 ρ3 ρ2 ρ0 δ2 δ1 µ1 µ2
ρ3 ρ3 ρ1 ρ0 ρ2 δ1 δ2 µ2 µ1
µ1 µ1 µ2 δ1 δ2 ρ0 ρ2 ρ1 ρ3
µ2 µ2 µ1 δ2 δ1 ρ2 ρ0 ρ3 ρ1
δ1 δ1 δ2 µ2 µ1 ρ3 ρ1 ρ0 ρ2
δ2 δ2 δ1 µ1 µ2 ρ1 ρ3 ρ2 ρ0

ρ1 H1 = {ρ1 ρ0 , ρ1 ρ2 } = {ρ1 , ρ3 } = ρ3 H1 ; H1 ρ1 = {ρ0 ρ1 , ρ2 ρ1 } = {ρ1 , ρ3 }


µ1 H1 = {µ1 , µ2 } = µ2 H1 ; δ1 H1 = {δ1 , δ2 } = δ2 H1

Thus, H1 is a normal subgroup and the blocked composition table shows


that it forms a group of these blocks as shown in Table 2.19.
————————————————–
Solution 2.20:
Given a, b ∈ G for an Abelian group we have ab = ba. Now s−1 as =
as−1 s = a for all a, s ∈ G. Hence each element of an Abelian group is conjugate
only to itself and thus each element separately is a class by itself.
———————————————————————————-
Solution 2.21:
Now Z3 × Z3 has 9 elements of product of (0, 1, 2)(0, 1, 2). Try the element
(1, 1) as a generator:

(1, 1) = (1, 1) ; 2(1, 1) = (1, 1) + (1 + 1 = (2 + 2)

3(1, 1) = (2, 2) + (1 + 1) = (0, 0)

Hence one obtains the identity (0, 0) of Z3 × Z3 only after three products
(additions). Now Z9 is a cyclic group of order 9. To be isomorphic to it, we
need an element in Z3 × Z3 which yields a 9-cycle. Since it is not so, Z3 × Z3
are not isomorphic to Z9 . We have just found another group of order 9; this
would thus be isomorphic to some other group of order 9.
————————————————–
Group Theory 61

Solution 2.22:
Given Z2 = {0, 1, }, Z4 = {0, 1, 2, 3} the elements of the group Z2 × Z4 are

{(0, 0), (0, 1), (0, 2), (0, 3), (1, 0), (1, 1), (1, 2), (1, 3)}

To check order of (0, 1) = (0, 1) ; 2(0, 1) = (0, 1) + (0, 1) = (0, 2)

3(0, 1) = (0, 2) + (0, 1) = (0, 3) ; 4(0, 1) = (0, 3) + (0, 1) = (0, 0)

So the order is four. Next the order of (1, 2),

(1, 2) = (1, 2) ; 2(1, 2) = (1, 2) + (1, 2) = (0, 0)

Hence, the order is two. The order of all the elements are given as

(1, 4, 2, 4, 2, 4, 2, 4), respectively.

The group is not cyclic as there are elements of order less than eight.
————————————————–
Solution 2.23:
Using the least common multiple; the largest orders are

(a) Z6 × Z8 = Z2.3 × Z2.2.2 , LCM is 2.3.2.2 = 24

(b) Z12 × Z15 = Z2.2.3 × Z3.5 , LCM is 2.2.3.5 = 60

————————————————–
Solution 2.24:
P
K = i∈I Ki , here k ∈ K. k = x1 + x2 + ...... + xm for some integer m
and xj belongs to some Kj . Let the order of xj be pj , and so,
(p1 , · · · pn )k = (p1 , · · · .pn )x1 + (p1 , · · · pn )x2 + · · · + (p1 · · · p2 )xn

= 0 + 0 + 0 + ···0 = 0

Hence, k is of order n and thus, K is a torsion group.


————————————————–
Solution 2.25:
Given (r ∈ R) assume that r + Z is of finite order.
Then m(r + Z) = Z for m a non-zero integer.
Next m(r + Z) = mr + Z and thus mr ∈ Z. Hence, r is a rational number.
Thus the torsion subgroup of R
Z
62 Group Theory in Particle, Nuclear, and Hadron Physics
Q
T(R
Z) ⊆ Z where Q is the subgroup of rational numbers.

n
However, if x = m then m, n ∈ z and m 6= 0.

n n
m(x + Z) = m( m + Z) = m( m )+Z =n+Z =Z
Q Q
Hence, x + Z is of finite order and thus Z ⊆ T(R R
Z ). Therefore T ( Z ) = Z.
————————————————–
Solution 2.26:
The highest order of elements in Z25 × Z9 is 25 × 9=225 and that of
Z5 × Z5 × Z9 is 5 × 9=45. Hence it not isomorphic.
————————————————–
Solution 2.27:
1089 = 32 × 112

so

3 3
11 11
33 33
(1) Z3 × Z3 × Z11 × Z11 :
→ Z33 × Z33

3 3
121
3 363
(2) Z3 × Z3 × Z121 :
→ Z3 × Z363

9
11 11
11 99
(3) Z9 × Z11 × Z11 :
→ Z9 × Z99

9
121
1089
(4) Z9 × Z121 :
→ Z1089
————————————————–
Solution 2.28:
Rewrite the product group as

Z × Z × Z × Z6 × Z10 × Z12
Group Theory 63

Clearly the Betti number is three. For the torsion number portion we take

2 2 4
3 3
5
2 2 60
Z6 × Z10 × Z12 = Z2 × Z3 × Z2 × Z5 × Z4 × Z3 →
→ Z2 × Z6 × Z60

And so the torsion coefficients are (2,6,60).


————————————————–
Solution 2.29:
< 1 >< 0 >< 0 > of {(0, 0, 0), (1, 0, 0), (2, 0, 0)}

Which has these elements. So the order of left cosets of this subgroup is
24
3 = 8.
————————————————–
Solution 2.30:
Z36 has 36 elements. The subgroup < 18 > has 2 members < 11 >=
{18, 0}. So the number of left cosets is 36
2 = 18.
————————————————–
Solution 2.31:
Z2 × Z4 or {0, 1} × {0, 1, 2, 3}. So has 2 × 4 = 8 elements

∼ {(0, 0), (0, 1), (0, 2), (0, 3)(1, 0), (1, 1)(1, 2), (1, 3)}

= {(0, 0), (0, 1), (1, 0), (1, 1)(0, 2), (0, 3)(1, 2), (1, 3)}

Left cosets of (1,2) are

e = (0, 0)+ < (1, 2) >= {(0, 0), (1, 2)} ; x = (1, 0)+ < (1, 2) >=
{(1, 0), (0, 2)}

y = (0, 1)+ < (1, 2) >= {(0, 1), (1, 3)} ; z = (1, 1)+ < (1, 2) >=
{(1, 1), (0, 3)}

The group structure is given in the following table

e x y z
{(0, 0), (1, 2)} {(1, 0), (0, 2)} {(0, 1), (1, 3)} {(1, 1), (0, 3)}
e{(0, 0), (1, 2)} {(0, 0), (1, 2)} {(1, 0), (0, 2)} {(0, 1), (1, 3)} {(1, 1), (0, 3)}
x{(1, 0), (0, 2)} {(1, 0), (0, 2)} {(0, 0), (1, 2)} {(1, 1), (0, 3)} {(0, 1), (1, 3)}
y{(0, 1), (1, 3)} {(0, 1), (1, 3)} {(1, 1), (0, 3)} {(1, 0), (0, 2)} {(0, 0), (0, 2)}
z{(1, 1), (0, 3)} {(1, 1), (0, 3)} {(0, 1), (1, 3)} {(0, 0), (1, 2)} {(1, 0), (0, 2)}
64 Group Theory in Particle, Nuclear, and Hadron Physics

So it does form a group and it is isomorphic to Z4 (one will have to


rearrange the order though).
————————————————–
Solution B.1:

X X
((AB)C)ij = (AB)ik Ckj = (Ail Blk )Ckj
k k,l
X
= Ail Blk Ckj
k,l
X X
(A(BC)ij = Aik (BC)kj = Aik (Bkl Clj )
k k,l
X
= Aik Bkl Clj
lk

As k and l are dummy indices, the above two are equal to one other.
————————————————–
Solution B.2:

detA = |A| = −5

   
1 −2 1 −3
Cofactor Ac = , tr(Ac ) =
−3 1 −2 1
so  
1 1 −3 −1
A =
5 −2 1
————————————————–
Solution B.3:
(a) let C=AB. then
X X X
tr(AB) = Cii = Aik Bki = Bki Aik
i i,k k,i

= T r(BA)
Note Tr(ABC)=Tr(BCA)=Tr(CAB)

(b)
X X ∼ ∼ ∼∼
^ = (AB)ji =
(AB) Ajk Bki = (B)ik (A)kj = (B A)ij
ij
k R
Group Theory 65
∼∼
so ^ = BA
(AB) ij

(c) X
((AB)∗ )ij = ((AB)ij )∗ = A∗il Blj

X
= (A∗ )(B ∗ )lj = (A∗ B ∗ )ij
l

so
(AB)∗ = A∗ B ∗

(d)

((AB)† )ij = ((AB)
^ )ij

^
= ((A∗B∗)
ij
∼ ∼
= (B ∗ A∗ )ij = (B ∗ A∗ )ij
so
(AB)† = B † A†
(e) We see this

(AB)(B −1 A−1 ) = A(BB −1 )A−1 = AIA−1 = I

so
B −1 A−1 = (AB)−1
(f )
(AB)† = B † A† = BA
so
(i(AB − BA))† = −i(B † A† − A† B † ) = +i(AB − BA)
————————————————–
Solution B.4:

X
−1
T rA0 = (S −1 AS)jj = Sjk Akl Slj
k,l
X
−1
= Slj Sjk Akl
k,l
X
δlk Akl = Akk = T rA
l
66 Group Theory in Particle, Nuclear, and Hadron Physics

————————————————–
Solution B.5:
   
0 1 0σ3 1σ3
σ1 ⊗ σ3 = σ3 =
1 0 1σ3 0σ3
 
0 0 1 0
 0 0 0 −1 
=
 1 0 0 0 

0 −1 0 0
————————————————–
Solution B.6:
From the definition of the direct product we see that

(A ⊗ B)† = A† ⊗ B †
as
A† A = AA† = In and B † B = BB † = Im
then

(A ⊗ B)(A ⊗ B)† = (A ⊗ B)(A† ⊗ B † )


= (AA† ) ⊗ (BB † )
= In ⊗ Im = Inm
and here A ⊗ B is unitary.

————————————————–
Chapter 3
Unitary Symmetry

3.1 Lie Groups and Lie Algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67


3.2 Some Specific Lie Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
3.2.1 SO(2): Special Orthogonal Group in Two Dimensions . . 73
3.2.2 SO(3): Special Orthogonal Group in Three Dimensions 76
3.2.3 SU(2): Special Unitary Group in Two Dimensions . . . . . 78
3.2.4 SU(3): The Special Unitary Group in Three Dimesions 80
3.3 Adjoint Representation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
3.4 Adjoint Representations of SU(2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
3.5 Cartan Subalgebra and Roots/Weight Space . . . . . . . . . . . . . . . . . . . . 87
3.6 Fractional Charges in SU(n) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
3.7 Appendix C: Linear Vector Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
3.8 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

3.1 Lie Groups and Lie Algebras


Lie groups are infinite groups which exhibit an additional property of their
multiplication laws being differentiable. One observes (see Appendix F) that
the differentiation merges well within the structure of a manifold. Thus we see
that a Lie group is a composite structure possessing both group properties as
well as manifold properties (see Appendix G).
Hence a Lie group is a differentiable manifold having in addition a group
structure, such that the group operation G×G → G and the map G → G with
g ∈ G given by g → g −1 are differentiable. If the dimension of the underlying
manifold is r, one says that G is an r-parameter Lie group [27], [25].
Let us commence with some examples of Lie groups.
General Linear Group: Given an n-dimensional space, let us transform
a point P continuously to another point P 0 while remaining in the same space.
This is a linear transformation specified by a n × n matrix which is invertible
(i.e., non-singular). This defines the general linear group GL(n). If the field of
numbers is complex C, then the group is GL(n, C). GL(n, R) is its subgroup
where the space is Rn . The number of parameters in GL(n, C) is 2n2 while
that in GL(n, R) is n2 .
If the determinant of the matrix is 1, then these groups reduce to SL(n, C)

67
68 Group Theory in Particle, Nuclear, and Hadron Physics
 
and SL(n, R) with the number of parameters being 2 n2 − 1 and n2 − 1 ,
respectively.
Unitary Groups: Given an n-dimensional vector in complex space C. ~r →
r~0 = R~r and ~r† → ~r0† = ~r† R† . Then ~r†~r = ~r† R† R~r conserves length →

r 2 = ~r+~r,
† † −1
while R R = 1 or R = R . So such a matrix is unitary and is denoted by
U (n). It is given by n × n invertible matrices. Note that U † U = U U † = 1.
This unitary group U (n) is of order n2 . If in addition det |R| = +1 then the
subgroup is denominated SU (n), the special unitary group in n-dimensions
and it is of order n2 − 1 .

————————————————–
Problem 3.1:
Show that SO(2) leaves x2 + y 2 invariant.
————————————————–
Problem 3.2:
For the SU (n) group defined in terms of n × n unitary matrices with unit
determinant, demonstrate that
(a) Any unitary matrix U can be written in terms of a Hermitian matrix
H as U = eiH
(b) T rH = 0
(c) Prove that (n − 1) number of generators are diagonal. (n − 1) is termed
the rank of the group.
————————————————–

Orthogonal Group
Given a vector in n-dimensional space in terms of z Euler angles. Let us
rotate around a specified origin. Then,

~r → ~r0 = R~r (3.1)

T
rT → r~0 = ~rT RT (3.2)
Thus if the radius squared is invariant,

~rT RT R~r = ~rT ~r (3.3)


Hence, RT R = 1, which implies RT = R−1 . This is called an orthogonal
matix and defines the orthogonal group O(n). If the elements are complex in
C, then the number of parameters in O(n) are n(n − 1). If the elements are
n(n−1)
in real space
TR, then the numbers of elements of O(n) are 2 .
As det R R = det |1|.
So det |R| = ±1.
With det |R| = +1 we obtain SO(n) the special orthogonal group in n-
dimensions with n(n−1)
2 parameters specifying it.
Unitary Symmetry 69

————————————————–
Problem 3.3: Show that the number of independent parameters required
to specify the groups O(n) and SO(n) are n(n−1)
2 for both.
————————————————–

Let us take n = 3. The group O(3) is larger as it contains elements with


both signs det R = ±1. Thus it consists of two disjoint sets with det R = +1
and det R = −1. The first one is denominated proper rotation SO(3) which
preserve the handedness of the space. The second one is termed improper
rotation which is a rotation coupled with space inversion.
 
−1
IS =  −1  ; IS X~ = −X ~ (3.4)
−1
IS and 1 in three dimensions form an Abelian group J. 1 and IS form
an Abelian group say J which is isomorphic to S2 the permutation group
in two dimensions. This is an invariant subgroup of O(3). Therefore each
element of O(3) is written as a direct product of proper rotation and J, i.e.,
O(3) = SO(3) ⊗ J.
Having specified a few often-encountered Lie groups in the domains of
particle, nuclear, and hadron physics, we now proceed to some in-depth study
of Lie groups.
Let us consider a Lie group G of linear transformation of order n. Then
each element of G depends on n-real parameters g(a1 , a2 ......an ) = g(aµ ) µ =
1, 2, 4 · · · n. We choose these elements such that the unit element of G corre-
sponds to the value

g(a1 = a2 = · · · an = 0) = e (3.5)
which is the identity element. Thus for the group G defined by a representation
Dn the unit element of G is represented as

Dn (g(aµ ))|aµ =0 = 1 (3.6)


with 1 being n × n the identity matrix.
We wish to study the infinitessimal transformation of the Lie group G in
the neighbourhood of the unit element, that is, the elements g(δa1 , δa2 , · · · δan )
corresponding to the infinitessimal values δa1 , δa2 , · · · δan of the parameters.
In all these (δaµ  1). So we can Taylor expand Dn (g(0 + δgµ )) as:

∂Dn (g(aµ ))
Dn (g(δgµ )) = 1 + δaµ |aµ =0 (3.7)
∂aµ
Let us define
∂Dn
Xµ = −i |a =0 (3.8)
∂aµ µ
70 Group Theory in Particle, Nuclear, and Hadron Physics

Dn (g(δaµ )) = 1 + iδaµ Xµ + · · · (3.9)


Xµ are termed generators of the group and −i are chosen to make these
Hermitian. The number of generators is equal to the number of parameters of
the group. Generators play a very basic role in Lie group theory [27], [25], as
we see below.
We can build a finite transformation from an infinite number of infinites-
a
simal transformations. So divide aµ into N units δaµ = Nµ and take the limit
N → ∞ as:

lim (1 + iδaµ Xµ )N = lim (1 + i Xµ )N = eiaµ Xµ (3.10)
N →∞ N →∞ N
In general there will be an infinite number of different sets of Xµ which
define a given Lie group. But we can define these within a similarity transfor-
mation. We choose a particular convenient set of Xµ . Thus as
P
Dn (aµ ) = ei µ aµ Xµ
(3.11)
And hence the entire representation of a Lie group is defined by the set of
generators Xµ , (µ = 1, ......n).
Now define a set of generator Xµ with parameter value aµ and another
with the parameter value bµ . Take the product of two elements of the group
eiaµ Xµ eibν Xν . Now as these form a group the product should be another ele-
ment of the same group as

eiaµ Xµ eibν Xν = eicα Xα (3.12)


Now we do not yet know how these quantum mechanical generators behave
when multiplied as matrices. Let us determine the right-hand side explicitly
by expanding it in a Taylor series and retaining up to second order terms.
Note that

icα Xα = ln eicα Xα


= ln(eiaµ Xµ eiβν Xν ) = ln(1 + A)


Where

A = eiaµ Xµ eiβν Xν − 1

1 1
= (1 + iaµ Xµ + (iaµ Xµ )2 + · · · ) · (1 + ibν Xν + (ibν Xν )2 + · · · ) − 1
2 2

1
A = i(aµ Xµ + bν Xν ) − aµ Xµ bν Xν − [(aµ Xµ )2 + (bν Xν )2 ]
2
Unitary Symmetry 71
2
Taylor expansion yields ln (1 + A) = A − A2 + .... and retaining only terms
up to the second order in a and b, we find,

A2 1
A− = i(aµ Xµ + bν Xν ) − [(aµ Xµ )(bν Xν ) − (bν Xν )(aµ Xµ )]
2 2
We define a commutator of two operators A and B as

[A, B] = AB − BA (3.13)
And thus,
1
icα Xα = i(aµ Xµ + bν Xν ) − [aµ Xµ , bν Xν ]
2
whence finally,
1
eiaµ Xµ eibν Xν = exp(i(aµ Xµ + bν Xν ) − [aµ Xµ , bν Xν ]) (3.14)
2
This is denominated the Baker-Campbell-Hausdorff formula, which is a
generalization of the normal exponential multiplication eA eB = eA+B which
in the above only works if the A and B operators commute [A, B] = 0 [28],[29].

————————————————–
Problem 3.4: Verify the validity of the Baker-Campbell-Hausdorff for-
mula up to the third order term.
 
iaµ Xµ ibν Xν 1
e e = exp i(aµ Xµ + bν Xν ) − [aµ Xµ , bν Xν ]
2
i

{[aµ Xµ , [aµ Xµ , bν Xν ]] + [bν Xν , [bν Xν , aµ Xµ ]]} (3.15)
12
————————————————–

Hence we see that the commutator [Xµ , Xν ] must be proportional to some


linear combination of the generator of the group, as this implies there exists
a closure within the generators. Thus,

[Xµ , Xν ] = iΣfµνλ Xλ (3.16)


λ

where fµνλ are termed the structure constants of the group. If these struc-
ture constants are known, then the commutation relation between all the
generators of the Lie group are completely specified. Hence the entire group
may be determined in a particular representation. The above algebra between
the generators of a particular Lie group is called the Lie algebra of the group
and is defined in terms of the structure constants.
It turns out that one can always associate a Lie algebra of order n to
72 Group Theory in Particle, Nuclear, and Hadron Physics

a Lie group of order n. One can actually prove that there is a one-to-one
correspondence between a Lie algebra and a simple connected Lie group. If it
emerges that two different groups happen to have the same Lie algebra (i.e.,
their Lie algebrae are isomorphic) then clearly there has to be a one-to-one
correspondence between the two groups – at least in the neighbourhood of
each element.
A Lie algebra of the kind that we found above in Equation 3.16 is a fun-
damental structure extending far beyond the Lie groups themselves. We thus
provide a more general definition of the same below.
A set of elements L is said to be a Lie algebra if it is a linear vector space
(see Appendix C) with a law of composition z = [x, y] where the right-hand
side is called the Lie bracket, such that

(a) If x, y ∈ L, then z ∈ L

(b) The Lie bracket is anti-symmetric [x, y] = − [y, z]

(c) It satisfies the Jacobi identity.

[x, [y, z]] + [y, [z, x]] + [z, [x, y]] = 0 (3.17)
Being a vector space, we choose a basis of independent element α1 , .....αn
so that
n
X
x= ai αi (3.18)
i=1

where ai are real coefficient. The number of independent elements n is the


order of the Lie algebra. The rank is defined as the number of commuting
elements in the Lie algebra.
An Abelian Lie algebra is one where all its elements commute [x, y] = 0.
L0 is said to be a subalgebra of L if L0 is a subset of the same Lie algebraic
structure. A subalgebra L0 of L is invariant if for each element x ∈ L0 and
y ∈ L, the Lie bracket [x, y] ∈ L0 . A simple Lie algebra does not contain
an invariant subalgebra, while a semi-simple Lie algebra does not contain an
invariant Abelian subalgebra. If a Lie algebra is semi-simple or simple, then
so are all the Lie groups corresponding to the Lie algebra.
Unitary Symmetry 73

X
2

r
0

0 x
n

FIGURE 3.1: Rotation of a two-dimensional vector – SO(2) group

3.2 Some Specific Lie Groups


3.2.1 SO(2): Special Orthogonal Group in Two Dimensions
Consider rotation in two dimensions. Rotate the vector ~r given in polar
coordinates as (r, α) to a vector ~r0 (r, α + θ). This is an active transformation
in contrast to a passive transformation, where a coordinate system may be ro-
tated in the opposite direction while keeping the vector stationary (see Figure
3.1).
Before rotation x1 = r cos α, x2 = r sin α and after rotation,

x01 = r cos (α + θ) = r cos α cos θ − r sin α sin θ = x1 cos θ − x2 sin θ

x02 = r sin (α + θ) = r sin α cos θ + r cos α sin θ = x1 sin θ + x2 cos θ

In matrix notation,

x01
    
cos θ − sin θ x1
= (3.19)
x02 sin θ cos θ x2
Given   0
x1 x1
X= , X0 =
x2 x02
74 Group Theory in Particle, Nuclear, and Hadron Physics

 
cos θ − sin θ
R (θ) = (3.20)
sin θ cos θ

X 0 = R (θ) X (3.21)
Note
  
cos θ1 − sin θ1 cos θ2 − sin θ2
R (θ1 ) R (θ2 ) = (3.22)
sin θ1 cos θ1 sin θ2 cos θ2

= R (θ1 + θ2 ) = R (θ2 + θ1 ) (3.23)


Hence the operator [R (θ1 ) , R (θ2 )] = 0 commutes.
Note that

AdjR (θ)
R−1 (θ) = (3.24)
det |R (θ)|
 
cos θ − sin θ
Finding the cofactor as R (θ) = .
sin θ cos θ
As AdjR (θ) is a transpose of the cofactor as
 
cos θ sin θ
R (θ) = (3.25)
− sin θ cos θ
So
 
−1 cos θ sin θ
R (θ) = = R (−θ) (3.26)
− sin θ cos θ
Note that R−1 (θ) = RT (θ). So it is an orthogonal matrix. Likewise,
θ = 0 yields the identity element. R (θ) forms a continuous group spec-
ified by a single parameter θ. This defines the rotation group in two di-
mensions, written as O (2). It is Abelian, R−1 (θ) = RT (θ) = R (−θ). So
RT (θ) R (θ) = R (θ) RT (θ) = 1.
Also as det R (θ) = 1, hence it is essentially SO(2), the special orthogonal
group in two dimensions.
Also,

     
1 −θ 1 0 1 0 −1
R (θ) → ≈ + iθ ≈ 1 + iθX (3.27)
θsmall θ 1 0 1 i 1 0

where
 
1 0 −1
X= (3.28)
i 1 0
is the generator of the group SO(2).
Unitary Symmetry 75

Note that in the mathematical literature, it is more common to define the


θ small expansion in terms of generators without the ‘i’ term. In physics, by
including the above ‘i’ term, we make the generator X as Hermitian and this
is the convention most physicists follow.

————————————————–
Problem 3.5: Prove that
d
R (θ) = eiθX where X = lim R (θ) (3.29)
θ→0 dθ
————————————————–

Let us now determine the irreducible representation of the group SO(2).


So define SO(2) on a finite dimensional vector space. As SO(2) is an Abelian
group all its irreducible representation (see Appendix D) are one-dimensional.
Thus given any vector |j > in a minimal invariant subspace under SO(3) we
demand that

X|j >= j|j > (3.30)


+iθα
R (θ) |j >= e |j > (3.31)
We label the state as j, which is a real number that is taken as the
eigenvalue of the Hermitian operator X. Clearly, R(θ1 )R(θ2 ) = R(θ1 + θ2 )
is satisfied by the above condition. In addition, to satisfy the global condition
R (θ) = R(θ ± 2π) one obtains constraints on the value of j. As e±iθ2πj = 1,
j is an integer.
Note for j = 0, R (θ) → U 0 = 1. This is the identity representation.
For j = 1, R (θ) ; R (θ) → U 1 (θ) = eiθ . Notice an isomorphism between the
SO(2) group element and real numbers on a unit circle in the complex plane.
Thus U 1 (θ) covers the unit circle once in the clockwise direction. For j = −1,
R (θ) → U −1 (θ) = e−iθ . Here the unit circle is covered counterclockwise.
Similarly for J = ±2, etc.
j ijθ
So the irreducible representation
  of SO(2) is given as R (θ) = e . Note
0 i
that the generator is X = . Let us diagonalize it. Its characteristic
−i 0
polynomial is λ2 − 1 = 0, and hence its√eigienvalues are λ± = ±1. Meanwhile,
its eigenvectors are e± = (∓e1 − ie2 ) / 2. Thus,

λe± = ±e± (3.32)

R (θ) e± = e∓iθ e± (3.33)


76 Group Theory in Particle, Nuclear, and Hadron Physics

3.2.2 SO(3): Special Orthogonal Group in Three Dimensions


We saw that the collection of rotations in two dimensions is an Abelian
group SO(2). But rotations in three dimensions form a non-Abelian group.
This is demonstrated by physically rotating an object, say a book, by π2 an-
gle, first in the sequence of rotations in x-, y- and z- directions, respectively.
Reversing the order of rotation sequence with the same angles does not return
the book to the original orientation.
An arbitrary rotation in three dimensions can be specified by a product
of at most three of 3 × 3 matrices given by rotations around axis 1, 2 and 3,
respectively (or equivalently in terms of the standard Euler angles).

   
1 0 0 cos θ2 0 sin θ2
R1 (θ1 ) = 0 cos θ1 − sin θ1  ; R2 (θ2 ) =  0 1 0 ;
0 sin θ1 cos θ1 − sin θ2 0 cos θ2
 
cos θ3 − sin θ3 0
R3 (θ3 ) =  sin θ3 cos θ3 0 (3.34)
0 0 1
These are real 3 × 3 unimodular (det R = +1) and orthogonal matrices
RT = R−1 .

————————————————–
Problem 3.6: Demonstrate that R1 , R2 and R3 are orthogonal, i.e., RT =
−1
R .
————————————————–

So these form a special orthogonal group in three dimensions SO(3). Con-


sider infinitessimal rotations,
3
X
Rj (θj ) = 1 + θj Xj (3.35)
j=1

Note that we are starting here with the mathematical convention. This
defines three generators of SO(3),

     
0 0 0 0 0 1 0 −1 0
X1 = 0 0 −1 ; X2 =  0 0 0 ; X3 = 1 0 0 (3.36)
0 1 0 −1 0 0 0 0 0

These are traceless and antisymmetric matrices


   
0 0 0 0 0 0
X1T = 0 0 1 = − 0 0 −1 = −X (3.37)
0 −1 0 0 1 0
Unitary Symmetry 77

(so anti-symmetric)
These obey the following algebra

[Xa , Xb ] = abc Xc (3.38)


As can be done for SO(2), one generalizes to SO(3) for rotations performed
for finite angles:
 
~ ~
R θ~ = eθ.X = eθ1 X1 +θ2 X2 +θ3 X3 (3.39)
 
where R θ~ are group elements of SO(3). As in SO(3) one generalizes to
visualize that here x2 + y 2 + z 2 = r2 is left invariant in SO(3).
So far the SO(3) generators are real. However, in quantum mechanics we
require Hermitian operators to make contact with quantum mechanics. Hence
let us define the generators as follows:

Jj = iXj (j = 1, 2, 3) (3.40)
So

     
0 0 0 0 0 i 0 −i 0
J1 = 0 0 −i ; J2 =  0 0 0 ; J3 =  i 0 0 (3.41)
0 i 0 −i 0 0 0 0 0

These are traceless and Hermitian matrices J † = J .




These obey the following albebra of SO(3):

[Ja , Jb ] = iabc Jc (a, b, c = 1, 2, 3) (3.42)


This is nothing but the angular momentum algebra. Hence the SO(3)
algebra is the same as the angular momentum algebra of quantum mechanics.
As demonstrated earlier, this is on account of the isotropy of space leading
to an invariance under rotations which in turn leads to the conservation of
angular momentum. It is a reflection of the same effect in quantum mechanics.
Indeed, this is a deep aspect of physical reality reflected in the Lie algebra of
SO(3).
We can redefine the algebra in terms of the raising and the lowering oper-
ators as

J ± = J1 ± iJ2 (3.43)
The Lie algebra of SO(3) can then be written as:

[J3 , J± ] = ±J± , [J+ , J− ] = 2J3 (3.44)


We now define a Casimir operator of SO(3) as
78 Group Theory in Particle, Nuclear, and Hadron Physics

J~2 = J12 + J22 + J32 (3.45)


= J± J± + J32 ± J3 (3.46)
A Casimir operator is defined so as to commute with all the generators of
the group:
h i
J~2 , Ji = [Jj Jj , Ji ] = Jj [Jj , Ji ] + [Jj , Ji ] Jj (3.47)

= Jj jik Jk + ijik Jk Jj (3.48)


= ijik (Jj Jk + Jk Jj ) = 0 (3.49)
This is because (Jj Jk + Jk Jj ) is symmetric under exchange of label j ↔ k
while jik is anti-symmetric under the same exchange.
The angular momentum Lie algebra above determines the eigenstates and
eigenvalues of the angular momentum operator in quantum mechanics. How-
ever the Casimir operator commuting with Ji above demonstrates that it can
commute with only one of the three generators. Conventionally we choose it
to be J3 . And thus both J~2 and J3 can be simultaneously diagonalized. If we
represent the state as |j, m >,

J~2 |j, m >= j (j + 1) |j, m > (3.50)


J3 |j, m >= m|j, m > (3.51)
The Hermiticity of J~ and that of J~2 ensures that j and m are real that
the eigenstate corresponding to different (j, m) states are orthogonal. In Ap-
pendices D and E, using the SO(3) Lie algebra commutation rules above, we
prove that (j, m) are both integers of half-integers.

3.2.3 SU(2): Special Unitary Group in Two Dimensions


As per our general definition above, the special unitary group SU (2) com-
prises rotations in two complex dimensions.
  We define the most general form of
α β
such a matrix, A ∈ SU (2) as A = given in terms of 8 real parameters
γ δ
(as each α, β, γ, δ is complex ). Because det A = 1,

αδ − βγ = 1 (3.52)
−1 †
Unitarity of the matrix U = U demands that
 ?
α γ?
  
δ −β
A−1 = = A† = (3.53)
−γ α β ? δ?
Thus
Unitary Symmetry 79

 
α β
A= (3.54)
−β ? α?
2 2
Let α = a1 + ib1 , β = a2 + ib2 , Note |α| + |β| = 1.
Thus ‘A’ has four independent parameters a1 , a2 , b1, b2 . In addition det A =
2 2
|α| +|β| = 1. This condition reduces the number of parameters to 3. One can
show that the matrix A forms a group with respect to matrix multiplication.
Now
? ?
   ? 
A† = A e? = α −β? =
α −β
(3.55)
β α β? α
And

T
α? β? α?
  
T ranspose (Cof actorA) −β
A−1 = = = (3.56)
det A −β α β? α

Hence A† = A−1 and thus the matrix is unitary.


So this is a special unitary group in two dimensions with three parameters
SU (2). We can define three generators (due to three independent parameters
of the A matrix),
   
∂A 0 1 ∂A 0 i
| α=1 = ; | α=1 = ;
∂Reβ −1 0 ∂Imβ i 0
β=0 β=0
 
∂A i 0
| = (3.57)
∂Imα α = 1 0 i
β=0
These are anti-Hermitian A† = −A and traceless 2 × 2 matrices. In turn,


these are related to Pauli matrices as


     
0 1 0 i i 0
→ iσ2 ; → iσ1 ; → iσ3 (3.58)
−1 0 i 0 0 i
     
0 1 0 i 1 0
σ1 = ; σ2 = ; σ3 = (3.59)
1 0 i 0 0 −1
These are the three generators of the SU(2) Lie algebra satisfying
hσ σ i σc
a b
, = iabc (3.60)
2 2 2
which is the same as the Lie algebra of SO(3) (J~ = ~σ ) 2

[Ja , Jb ] = iabc Jc (3.61)


80 Group Theory in Particle, Nuclear, and Hadron Physics

This is the angular momentum Lie algebra of SO(3). Thus the Lie algebras
of SU(2) and SO(3) groups are isomorphic.
Hence given the fact that the SU (2) and SO(3) Lie algebras are isomor-
phic, what does it say about the groups themselves? Note that for SO(3),

R = eiθi Ji (i = 1, 2, 3) (3.62)
And for SU (2),
iφj σj
U =e 2 (j = 1, 2, 3) (3.63)
φ
Globally the behavior of θi and 2j is the same. When θi = 2π the group
φ
element in SO(3) is the same as 2j = π the element of SU (2) where SU (2) →
φj
−1. When θi = 4π then 2 = 2π leads to +1 for SU (2). So there is a 2-1
mapping of elements of SU (2) onto those of SO(3). In particular,
 
1 0  
& 1 0 0
SU (2) : 0 1  0 1 0 : SO(3) (3.64)
1 0
− % 0 0 1
0 1
So though the Lie algebrae of SU (2) and SO(3) are isomorphic, the cor-
responding Lie groups of  SU (2)
 and SO(3) are homomorphic.
1 0
We may add σ0 = to the Pauli matrices σ1 , i = 1, 2, 3 to have
0 1
all the generators of the unitary group U (2). As σ0 commutes with all σi , it
can stand for a conserved quantum number like a baryon number or a lepton
number. Note that the group U (1) ⊗ SU (2) is not semi-simple and it is a
direct product of the Abelian group U (1) with the simple group SU (2). Note
that the groups U (2) and U (1)⊗SU (2) are locally isomorphic and hence have
the same algebra.
There is a two-to-one homomorphism between U (1) ⊗ SU (2) and U (2) as
both the elements:
   
iθ 1 0 i(θ+π) −1 0
e and e (3.65)
0 1 0 −1
 iθ 
e 0
of U (1) ⊗ SU (2) correspond to element of U (2).
0 eiθ
For the next unitary group a similar correspondence holds for U (1)⊗SU (3)
and U (3) or we may generalize this to the relationship between U (1) ⊗ SU (n)
and U (n).

3.2.4 SU(3): The Special Unitary Group in Three Dimesions


We define SU (3) for the so-called SU (3)F flavour group. We can gener-
alize this language for any other SU (3) group, like the colour of the strong
Unitary Symmetry 81

interaction. Let the fundamental representation, ( np ) of the SU (2) group be


p
generalized to three entities n ,
λ
let us relabel the vector with new unknown entities as
 
u
ψ= d  (3.66)
s
Then the group SU (3) is represented by the transformation
2
  nX −1
0 1 ~~
ψ = U ψ , U = exp iθ.λ = exp {i θa Fa } (3.67)
2 a=1

where λ are 8 traceless Hermitian matrices. Remember SU (3) is specified by


32 − 1 = 8 traceless Hermitian matrixes. Out of these 3 − 1 = 2 are diagonal
generators. These 8 matrices are generalizations of ~σ of SU(2). These in the
standard notation are

     
0 1 0 0 −i 0 1 0 0
λ1 =  1 0 0  ; λ2 =  i 0 0  ; λ3 =  0 −1 0 ;
0 0 0 0 0 0 0 0 0
     
0 0
1 0 0 −i 0 0 0
λ4 =  0 0
0  ; λ5 = 0 0 0 ; λ6 = 0 0 1 ;
1 0
0 i 0 0 0 1 0
   
0 0 0 1 0 0
1
λ7 =  0 0 −i  ; λ8 = √  0 0 1  (3.68)
0 i 0 3 0 1 0
Note that λ3 , λ8 are diagonal. Also note how SU (2) subgroups are con-
tained within SU (3). λ1,2 exhibit the structure:
 
0
 σ1,2 0  (3.69)
00 0
Leading to SU (2)I the isospin subgroup acting on ( vd ) vector. Now λ6,7 is
 
0 0
 0  (3.70)
σ1,2
0
This is a sub-group called U-spin acting on ( ds ) vector. Meanwhile λ4,5 is
related to the third SU (2)V , V-spin acting on ( us ) vector.
Hence, there are three subgroups called the I-spin, the U-spin and the
V-spin, acting on ( ud ), ( ds ) and ( us ) vectors, respectively.
82 Group Theory in Particle, Nuclear, and Hadron Physics

The isospin operator of SU (3) is defined as:


 
1 0 0
1 1
F3 = λ3 = 0 −1 0 (3.71)
2 2
0 0 0
So F3 acting on the fundamental representation yields isospin eigenvalues
as:
      1 
u 1 0 0 u +2u
1
F3 d = 0 −1 0 d = − 21 d (3.72)
2
s 0 0 0 s 0
Next we define the hypercharge operator as:
2 2 1 1
Y = √ F8 = √ λ8 = √ λ8 (3.73)
3 3 2 3
And
      1 
u
1
1 0 0 u 3u
Y = d = 0 1 0  d =  13 d  (3.74)
3
s 0 0 −2 s − 23 s
These two define the two quantum numbers of the fundamental states.
One writes the SU (3) Lie algebra as:
 
1 1 1
λi , λj = ifijk ( λk ) (3.75)
2 2 2
As Fi = 12 λi ,

[Fi , Fj ] = ifijk (i, j, k = 1, 2, 3) (3.76)


fijk are the structure constants of SU(3). fijk are anti-symmetric under
the exchange of any pair of indices i, j, k. The electric charge is Q = F3 + Y2 .
Hence
   2 
u 3u
Q d = − 13 d (3.77)
s − 13 s

——————————————-
Problem 3.7: Determine the non-zero structure constraints of SU (3).
————————————————–

The Casimir operator for SU (2) is I~2 and it commutes with all the gener-
ators of SU (2), [I~2 , Ii ] = 0. If
Unitary Symmetry 83

1
C = I~2 = {I+ I− + I− I+ } + J32
2
1
= {I+ , I− } + J32 (3.78)
2
In SU (3) the invariant operator is
8
X
F~ 2 = Fi Fi
L=1

1 1 1
= {I+ , I− } + I32 + {U+ , U− } + {V+ , V− } + F82
2 2 2
where I± = F1 ± iF2 , I3 = F3 , U± = F6 ± iF7
2
V± = F4 ± iF5 , Y = √ F8 (3.79)
3
And the Casimir operator commutes with all the generators of SU (3)

[F~ 2 , Fi ] = 0, i = 1, 2.......8 (3.80)


Note the significance of the raising and lowering operators like I± , U± , V±
is identical to what we found for the corresponding analogous situation for
the irreducible representation of SO(3).

————————————————–
Problem3.8: In SU(3) we saw the eighth generator defined as diagonal:
1 0 0
λ8 = √13 0 1 0 
0 0 −2
Call + √13 as the normalization factor. In SU (2) we take the normaliza-
th
tion factor as +1. Guess the normalization factor for the diagonal (n2 − 1)
generator of SU (n). So what is it for SU (4) and SU (5)?
————————————————–

~ U
There are three SU (2) subgroups I, ~ and V
~ of SU (2). We define Q =
Y
F3 +
h→ 2 . Let us define Z = Q − Y One finds
− i
T ,Y = 0
h→
− i
U,Q = 0
h→
− i
V ,Z = 0

————————————————–
Unsolved Problem 3.1 : Confirm the above three commutation relations.
————————————————–
84 Group Theory in Particle, Nuclear, and Hadron Physics

Note that the lowest dimensional representation, i.e., three-dimensional in


the case of SU (3), is called the fundamental representation. Hence quarks
belong to the fundamental representation of SU (3). Another representation
which has the same dimension as the number of generators of SU (3), i.e.,
8-dimensional representation, is called the adjoint representation.

3.3 Adjoint Representation


We defined the Lie algebra associated with a particular Lie group as:
X
[Xλ , Xµ ] = i fλµν λν (3.81)
ν

where the generators Xρ are Hermitian and the structure constants are real.
Moreover, fλµν = −fµλν due to the property of the commutator [ , ].
Now these generators, in order to be called a Lie algebra, should addition-
ally satisfy the following Jacobi identity:

[Xλ , [Xµ , Xν ]] + [Xµ , [Xν , Xλ ]] + [Xν , [Xλ , Xµ ]] = 0 (3.82)


By expanding we can see that, independent of the structure constants and
only due to the definition of the property of commutation, the above Jacobi
identity is trivially satisfied. Hence as such it does not seem to provide any new
information within the structure of the Lie algebra. Thus, the Jacobi identity
is trivial if the left-hand side of the Lie algebra is utilised. But interestingly,
new information is gleaned when the right-hand side of the Lie algebra is used
in the Jacobi identity. Substituting Equation 3.81 in 3.82
P
i τ {[Xλ , fµντ Xτ ] + [Xµ , fνλτ Xτ ] + [Xν , fλµτ Xτ ]} = 0

= i2
P
τ,σ (fλτ σ fµντ + fµτ σ fνλτ + fντ σ fλµτ ) Xσ = 0

We assume that the generators Xσ form a basis of the Lie algebra


P
τ fλτ σ fµντ + fµτ σ fνλτ + fντ σ fλµτ = 0

Let us swap a few indices,


X
fλτ σ fµντ − fµτ σ fλντ − fτ νσ fλµτ = 0 (3.83)
τ

Again trivially satisfied. But now define an n-parameter set of n numbers


of n × n matrices {Ta : a = 1, 2, ....n} as the adjoint representation, defined in
terms of a structure constant,
Unitary Symmetry 85

(Tα )βγ = −ifαβγ (3.84)


Then writing the above expression connecting the structure constants in
Equation 3.81 in terms of the components of the adjoint representation, we
obtain:
P 
τ − (Tλ )τ σ (Tµ )ντ + (Tµ )τ σ (Tλ )ντ − i (Tτ )νσ fλµτ = 0
P 
τ − (Tµ )ντ (Tλ )τ σ + (Tλ )ντ (Tµ )τ σ − i (Tτ )νσ fλµτ = 0

We have here the (ν, σ) element of a matrix equation


X
Tλ Tµ − Tµ Tλ − i fλµτ Tτ = 0
τ

So
X
[Tλ , Tµ ] = i fλµτ Tτ (3.85)
τ

Now we see that the above adjoint representation Lie algebra is isomorphic
to the Lie algebra, but it is a different representation for which it holds, i.e.,
of adjoint representation Equation 3.84. Hence let us conclude that the Jacobi
identity is nothing but the adjoint representation Lie algebra “in disguise”!
This adjoint representation is a primary entity as an n-dimensional vector
built upon n-generators of the group, similiar to a self-representation. Let us
choose the basis in which the adjoint representation satisfies:

T r(Tτ Tν ) = a δτ ν (3.86)
Where ‘a’ is a scalar. Substituting for Tτ we obtain:
 
1
Tr P [Tλ , Tµ ]Tν = aδντ
i τ fλµτ
So
1 X
T r {[Tλ, Tµ ] Tν } = fλµτ δντ
ia τ

So fλµν = − ai Tr {Tλ, Tµ } Tν
Now the trace (see Appendix B) does not change under cyclic permuta-
tions and is negative under anti-cyclic permutations. We see that

fλµν → T r {[Tλ, Tµ ] Tν } = T r {Tλ Tµ Tν − Tµ Tλ Tν }

= T r {Tν Tλ Tµ − Tλ Tν Tµ }
86 Group Theory in Particle, Nuclear, and Hadron Physics

= T r {[Tν , Tλ ] Tµ } = −T r {[Tλ , Tν ] Tµ } = −fλνµ

Thus we find that fλνµ is a completely antisymmetric tensor:

fλµν = fµνλ = fνλµ = −fλνµ = −fµλν = −fνµλ (3.87)

————————————————–
Unsolved Problem 3.2: Confirm all the expressions in Equation 3.87
above.
————————————————–
Problem 3.9: Discuss how structure constants behave under exchange of
indices in the groups SU (2), SU (3) and SU (n) for n > 3.
————————————————–

Note that the matrices Tα provide a representation which is irreducible.


Moreover, observe that all the Tα are antisymmetric and hence none of these
are diagonal. We do need to know as to how many generators can be placed in
a diagonal form in a particular group. This is achieved by means of a unitary
transformation, U Tα U † = Fα . These yield equivalent representations in which
we ensure that the maximum number of generators are diagonal. Now we know
that it is one, (T3 ), in SU (2) and two, (λ3 , λ8 ), in SU (3). One and two are
the ranks of the SU (2) and SU (3) algebras, respectively. The rank of the Lie
algebra always informs us as to how many generators of the Lie algebra can
be put in a diagonal form.

3.4 Adjoint Representations of SU(2)


As we saw in our discussion on adjoint representations, the Jacobi identity
is actually an independent Lie algebra in disguise. The structure constraints
themselves specify this algebra.
Let us take SU(2) as specifying the isospin group where ( np ) stand for the
fundamental representation of SU(2), the isospin group (see Chapter 11 for
details). Its Lie algebra is

[Ji , Jj ] = iijk Jk (3.88)


Take

(Jk )ij = −iijk (3.89)


This is the matrix specified by the structure constants of the SU (2) group.
As we saw above the structure constants, on account of the Jacobi identity,
Unitary Symmetry 87

satisfies a Lie algebra isomorphic to the defining Lie algebra. The set of matri-
ces specifying the Lie algebra represent the isospin I=1 state with the matrices
(generators) given as

     
0 0 0 0 0 −i 0 −i 0
I1 = 0 0 −i ; I2 = 0 0 0  ; I3 =  i 0 0 (3.90)
0 i 0 i 0 0 0 0 0

This specifies the Cartesian form of the pion field →



π = (π1 , π2 , π3 ), i.e.,

Ii |πj >= iijk |πk > (3.91)


These pions πi (i = 1, 2, 3) are not eigenstates of the charge operator and
hence this representaton is not realized physically. This is also due to the fact
that the matrices ti are antisymmetric, so none is diagonal. We perform a
unitary transformation to bring it to a diagonal form to obtain physical pion
states in a spherical basis:

−1 1
|π + >= √ {|π1 > +i|π2 >} , |π − >= √ {|π1 > −i|π2 >} , |π 0 >= |π3 >
2 2
(3.92)
These have proper charges for the pion triplet.

————————————————–
Problem 3.10: Determine the basis in which the isospin operators I~2 and
I3 are diagonal.
————————————————–
Unsolved Problem 3.3: Determine all the diagonal matrices I1 , I2 and
I3 through explicit calculations of < πµ |Ii | πν > matrix elements.
————————————————–

3.5 Cartan Subalgebra and Roots/Weight Space


Given a Lie group with n-generators, we say it is an n-dimensional group.
First we try to diagonalize as many of these n-generators as possible. Let us
take linear combinations of these generators and call these Hi , i = 1, z.....m
where m < n, i.e., m the number of Hi is less than the total number of
generators. Thus,
X
Hi = aiα Xα
α
88 Group Theory in Particle, Nuclear, and Hadron Physics

[Hi , Hj ] = 0
These maximal number of commuting generators are called the Cartan
subalgebra and these generators are denominated Cartan generators [27],
[29], [30]. Here m is termed the rank of the algebra.
Now we know that the Cartan generators are linear combinations of the
diagonal generators of the group. And hence for the SU (3) group, strictly
speaking, a Cartan subalgebra would be Hi = aλ3 + bλ8 (i = 1, 2), λ3 and
λ8 are the diagonal generators of SU (3) defined earlier. One normally chooses
H1 = λ3 (a = 1, b = 0), and H2 = λ8 (a = 0, b = 1). As we see this
convenient choice leads to defining the quarks properly as corresponding to
the fundamental representation. However for the adjoint representation one
has to be careful, as we shall see in Chapter 7.
Note that the non-Cartan generators are called Ei (i = 1, 2....n − m). So
in SU(2) : H1 = J3 and E1 = J1 , E2 = J2
For a typical larger Lie group where m-dimensional quantum numbers
{h1 , h2 .......hm } of the Hi (i = 1, 2....m) Cartan subalgebra may distinguish
the states by labelling them as such. In addition the Lie group would have
a number of Casimir operators (like J 2 of SU (2)) whose number would be
equal to the rank of the group. Let us denote these Casimir operators as
Cα , (α = 1, 2, . . . m). In addition, is it possible that there are other indepen-
dent functions of the generators, let us call them Sβ , which commute with the
complete set of Hi , but not with the complete set of Xj ? The number β will
depend upon a particular group.
Now the sets Hi , Cµ and Sj are mutually commuting operators. Hence
these can be diagonalized simultaneously. Thus one can find a set of basis
vectors which are eigenstates of all these operators simultaneously. Hence let
us label them as c = {c1 , c2 .....cm } ,h = {h1 , h2 .....hm } and s = {s1 .........sβ }:
Cα Ψchs = cα Ψchs
Hi Ψchs = hi Ψchs
Sβ Ψchs = sβ Ψchs


So for SO(3), C1 = J 2 and H1 = J3 , whence there are no Sk (this is as
per the conventional understanding). Thus these Eigenstates are specified by
the two eigenvalues j, m or the state is written as |j, m >. However in larger
groups, the eigenstates may not be uniquely specified by the states of c and h,
and then one would invoke sk to distinguish between the different eigenstates.
Note that the above Cartan subalgebra fixes only m-number of genera-
tors Hi . Therefore, there are (n − m) number of generators left which do not
commute with Hi (i = 1, .....m). Cartan however showed that these other gen-
erators can always be put in the form of step operators like J± in SO(3) or
SU (2). So the remaining generators can be linearly transformed in a set of
operators Ea , a = 1, 2, . . . (n − m), which exhibit the following commutation
with Hi ,
Unitary Symmetry 89

[Hi , Ea ] = kai Ea (3.93)


Applying this to eigenstates Ψchs we get,
Hi {Ea Ψchs } = (hi + kai ) {Ea Ψchs }
Thus {Ea Ψchs } is an eigenstate of Hi with eigenvalue (h + kai ). Hence
Ea Ψchs should be a multiple of state Ψc(h+ka )s and so these are step-up op-
erators of the Lie group. The set of (n − m) vectors ka = {ka1 , ka2 ......kam }
are called weight vectors of the group. These values themselves are termed
the weights. For SO(3) rank m = 1, so the weight vector is one-dimensional
with a weight k = ±1. For the adjoint representation, to signify its special
character with respect to the other representations (states), the word ‘weight’
is replaced by the word “root”. Similarly we can derive step-down operators
as well. And analogous to the angular momentum algebra one can obtain all
the weights of the representation.
In general an irreducible representation of the Lie group is identified by
eigenvalues of the Casimir operator c = {c1 , c2 .....cm } and may be written as
< Ψchs |Xλ | Ψchs >.
For the adjoint representation, as the dimensionality is fixed, the state will
be satisfied by the eigenvalues of the commuting Hi and Sk . Remember that
for the adjoint representation, matrices are antisymmetric. Hence the unitary
transformation has to be performed to obtain an equivalent representation in
which Hi and Sk are diagonal.
Let us discuss the root-vector and the weight-vector of the SU (2) to ap-
preciate the similarity and the differences between the two. We saw that the
generators (Ik )ij = −iijk for SU(2) as antisymmetric matrices. As per the
Cartan criterion we have to diagonalize the maximum number of generators
of the Lie group. We solved this as a problem earlier and noticed that only
one generator is
 
1 0 0
H1 = 0 0 0 
0 0 −1
and two more generators assosicated with the non-Cartan generators are
   
0 1 0 0 −i 0
1  1
E1 = √ 1 0 1 ; E2 = √  i 0 −i
2 0 1 0 2 0 i 0
This is true for the adjoint representation and hence there are three eigen-
vectors.
     
1 0 0
v1 = 0 ; v2 = 1 ; v3 = 0
0 0 1
90 Group Theory in Particle, Nuclear, and Hadron Physics

kk + 1 k + 0 k = +1
1
3 2

FIGURE 3.2: Root-vector representation of the adjoint representation of


SU (2)

which is the same as the number of generators for the adjoint representation.
It is a one-representation as rank m = 1. Thus, the root vectors are given as:

k1 = +1; k2 = 0; k3 = −1
We draw the root vector diagram for the adjoint representation of SU (2)
in Figure 3.2.
Note that there are three generators, three eigenvectors and three root-
vectors for SU (2). Remember that for the adjoint representation of a Lie
group, the number of generators, eigenvectors and root-vectors are the same.
Let us see how do the weight vectors arise in SU (2). We know that the state
Ψj with j integer or half-integer arises naturally (see Appendix E). However,
the weight-vectors perpetually appear the same as in Figure 3.2, as the raising
and lowering operators will always change it by ±1.
However, the physical state space will change for each representation. For
j = 12 , 1, 32 , . . . there are only two, three and four physical states, respectively.
We plot the graphs for these weight spaces of SU (2) in Figure 3.3. Note the
dark dots which represent a particular weight of the respective weight-vectors.
Comparing Figure 3.2 and Figure 3,3, one notes the similarity and the
difference for the state ~j = 1 and the adjoint representation ~k = 1. What is
common is that both of them have the same size, i.e., both these are vectors
of magniturde ‘1’, but the difference between them is indicated by the fact
that the first is representaed by an arrow and the second by dots.
Note that each of these eigenvectors has their own eigenvalues of H1 . Hence
Unitary Symmetry 91

2 2 2

1 1 1

O O O

-1 -1 -1

-2 -2 -2

_!_ 3_
j j = 1 J
2 2

FIGURE 3.3: Weight vectors of the j-representation of SU (2)

each of these diagrams indicates that the eigenvalue changes by one unit in
each step moving from j to −j. In Figure 3.3 these diagrams correspond
to the eigenvalues of the eigenvectors and therefore each representation has
a different number of dots. However, each representation always arises from
three generators H1 , E+1 , E−1 . So either they do not move at all (H1 ); or they
move up (E+1 ) by one unit; or they move down (E− ) by one unit.
Therefore, the root vector of the adoint representation is special. The size
of its root vector, the size of its eigenvector and the number of generators
is the same, i.e., three. This renders it a unique representation. The arrow
in Figure 3.2 signifies that uniqueness. So the adjoint representation with
~k = 1 and the other ~j = 1 representation which is built differently, as from
the combination of two fundamental representations of dimension two, are
mathematically quite different! And there is no basic reason as to why this
fundamental difference in mathematics would not demand differences in the
corresponding physical realm as well.
Thus for SU(2)I the adjoint representation is given in Figure 3.2 for ~k = 1
and is used to define the physical pion as

1 1
|π + >= − √ {|π1 > +i|π2 >} , |π − >= √ {|π1 > −i|π2 >} , |π 0 >= |π3 >
2 2
(3.94)
So the adjoint representation ensures correspondences with the charges
of the physical pion. Remember that the Cartesian components (π1 , π2 , π3 )
of the vector representation of the group SU (2) do not correspond to the
92 Group Theory in Particle, Nuclear, and Hadron Physics

1 Y
n P

+
Z' zo E

_!_ 0
-1 +- +1
~2 ^2

o
-1

FIGURE 3.4: Weights of the baryon octet states in SU (3)

physical pion. Hence the arrows in Figure 3.2 signify associations with the
physical reality of the adjoint representations. The charge is obtained from
complexification as a global effect. We emphasise the important role that
the electric charge of pions displays here. It exhibits that the pion in the
Cartesian components form is not an eigenstate of the electric charge, while
in the spherical polar component form, the pion is indeed an eigenstate of the
elctric charge operator. What connects these two coordinate sysytems? Just a
global geometric transformation of complexification. Hence the electric charge
operator here is a basic and global entity which has no components or is not
made of other parts. Thus the charge of the pion as an adjoint representation
above exhibits this property.
However as per the Fermi-Yang model [31], pions can be treated as com-
posites of nucleon-antinucleon complexes. In the Fermi-Yang model one takes
N = ( np ) as a fundamental representation SU (2) and constructs the mesons
as N N̄ states. We associate pions as:
1
π + = pn̄, π − = p̄n, π 0 = √ (−p̄p + n̄n) (3.95)
2
The charges above are canonically defined as for the SU (2)-isospin group
with T3 as the isospin quantum number and B is the baryon number:
B
Q = T3 + (3.96)
2
Now as far as charges go, these are the proper pions made up of N N̄ .
But these pions are composite entities. Also the charge operator itself is a
Unitary Symmetry 93

composite operator. Hence though they correspond to the same pions, the
picture is entirely different. We may associate these with correspondence to the
dots in Figure 3.3 for I=1. Also, in contrast to the global nature of the charges
of the adjoint representation pions, here the charges are “local” in nature. Thus
there is a Manichaean duality in the physical description of the pions in the
group SU (2)I . This is the physical explanation for the mathematical duality
inherent in the ~k = 1 (Figure 3.2) and J~ = 1 (Figure 3.3) pictures. Indeed,
this is the origin of the “arrow” and the “dot” difference in these two figures.
But remember that it is the global and non-composite nature of the definition
of the electric charge associated with the adjoint representation of the SU(2)
group that has permitted the ability to perceive the above duality.
Now for the group SU (3) we saw that, taking the Cartan subalgebra to
correspond to the diagonal matrices T3 = √12 λ3 and Y = √13 λ8 , we can define
the fundamental representation with (u, d, s)-quarks as eigenstates of (T3 , Y )
operators. Now baryon states, as we shall show in Chapter 7, are

3 ⊗ 3 ⊗ 3 = 1 ⊕ 8 ⊕ 8 ⊕ 10
This also yields an 8-dimesionial octet representation . This is an extension
of Figure 3.3 for SU (2) and is given in Figure 3.4. Note here the dots as
generalized for SU (3) from the corresponding SU (2) figure. In contrast to the
corresponding 8-dimesional adjoint representation of SU (3), the generalization
of Figure 3.2 of SU (2) produces Figure 3.5. Note the arrows here for the adjoint
representation.
In Figure 3.4 the eigenvalues are called the weights and in Figure 3.5
these are termed roots. It is important to distinguish the more basic nature
of the roots vis-a-vis the weights [[29] page 115-116] here. Note that the roots
only stand for the “lone” adjoint representation. There seems to be a dual
description of the same mathematical reality and this should manifest itself in
terms of a dual phyiscal reality also. Its emergence in SU(3) shall be discussed
in Chapter 7.

3.6 Fractional Charges in SU(n)


We have seen that for the SU (3) group the electric charges defined as Q =
T3 + Y2 have the values 23 , − 13 and - 31 for the u-, d- and s-quarks, respectively.
This was a group theoretical success of the SU (3) model for providing the
correct (i.e., experimentally determined) charges of the quarks in terms of the
two diagonal generators of SU(3). Subsequently new flavour c-, b- and later
t- quarks, corresponding to the groups SU(4), SU(5), SU(6), were discovered.
94 Group Theory in Particle, Nuclear, and Hadron Physics

1 Y
n P

Z"
z° z+
0 T
•1 ^
2 + _!_ +1 3
7

o
-1

FIGURE 3.5: Roots of the baryon octet states in SU (3)

So what does the group predict for the electric charges of c-, b- and t-quarks?
The point is, if SU (3) is so successful in accounting for the charges of u-, d-
and s- quarks (as belonging to its fundamental representation) so we should
be able to predict the electric charges of the quark of these belonging to say
the SU (4), SU (5) and SU (6) groups, respectively.
However the above conjecture, while mathematically valid, turns out to
be a disaster when applied to 4-, 5- and 6-flavours within SU (4), SU (5) and
SU (6) groups. Let us try to comprehend this problem.
First we define nine traceless matrices in SU (3) as:
1
(Aij )µν = δiν δjµ − δij δµν , (i, j, µ, ν = 1, 2, 3) (3.97)
3
Only eight of these matrices are independent as

A11 + A22 + A33 = 0 (3.98)


These Aij are related to the generators of SU (3) in the standard form given
above. Here we are interested in those Aij which are related to Q, Y and Z (as
defined in SU (3) above).
1
A11 = δ1ν δ1µ − δ11 δµν (3.99)
3
And thus,
Unitary Symmetry 95

1 2 1 1
(A11 )µ=1,ν=1 = 1 − = , (A11 )µ=2,ν=2 = − , (A11 )µ=3,ν=3 = − ,
3 3 3 3
and hence for the electric charge Q we have
 2 
3 0 0
A11 = Q =  0 − 13 0  (3.100)
0 0 − 31
The other terms are
1
 
3 0 0
A22 = −Z = −  0 − 23 0  (3.101)
1
0 0 3
 1 
3 0 0
A33 = −Y = −  0 1
3 0  (3.102)
0 0 − 32
Thus Aij yields diagonal matrices which are related to the invariants of
different SU (2) subgroups. Most important here is A11 = Q, that is the electric
charge of quarks in the fundamental representation of SU (3).
With the discovery of charm as the fourth quark, one was forced to re-
sort to groups larger than SU (3). Especially favoured were the ones which
would permit the existence of three conserved quantum numbers, isospin, hy-
percharge and the new C-number for the charm quark. A priori there were
several groups which were possible candidates. However, the simplest exten-
sion to SU (4) was the one which was found to work the best in fitting the
experimental data.
Let us define the 42 = 16 traceless matrices Aij of SU(4) as
i
(Aj )µν , (i, j, µ, ν = 1, 2, 3, 4),
1
(Aij )µν = δiν δjµ − δij δµν (3.103)
4
We find
3 
4 0
− 41
A11 = 
 
 (3.104)
 − 14 
0 − 41
plus

− 14 − 41
    1 
0 0 4
0
3
− 14 − 14
A22 = 
  3   4  
4  , A3 =   , A4 =  
 − 41   3
4
  − 14 
0 − 14 0 − 41 0 3
4
96 Group Theory in Particle, Nuclear, and Hadron Physics

Now generalizing from the group SU (3), the group theoretical prediction is
that A11 should be equal to the electric charge for the SU (4) (u,d,s,c)-quarks.
This is a clear-cut prediction for quark charges in SU (4). What about
empirical results? These reject the above group theoretical prediction. Exper-
imentally u-, d-, s- charges remain 32 ,- 13 and - 13 (as was true in SU (3)) and the
c-quark charge is 23 . This was a gross failure of the SU (4) group theoretical
idea. So does it mean that the relevant group with the c-quark is not SU (4)
but some other group?
However, as a redeeming feature, the classification of states as per SU (4)
work well and we believe that the correct group to include c-quark is indeed
SU (4). However, what is not correct is the SU (4) prediction for the electric
charges. So how do we obtain the correct electric charges in SU (4)?
In SU (4) the third diagonal generator is given as (see the problem above):
 
1 0 0 0
1 0 1 0 0 
λ15 = √   (3.105)
6 0 0 1 0 
0 0 0 −3
The so-called Gell-Mann-Nishijima relation of SU (3) is
Y B+S
Q = T3 + = T3 + (3.106)
2 2
To account for the experimentally confirmed charge of Qc = 23 we expand
to (not to A11 of SU (4)) as a generalized Gell-Mann-Nishijima expression,
B+S+C
Q = T3 + (3.107)
2
where T3 is + 21 and - 12 for u- and d-quark, respectively, and zero for all others.
S is zero for all except -1 for the s-quark and C is +1 for the c-quark and zero
for all others.
Note that in SU (3),

1 1 1 1 1
T3 = λ3 , B = λ0 , S = √ λ8 − λ0 , Y = B + S = √ λ8 (3.108)
2 3 3 3 3

1 1 1
Q= λ3 + √ λ8 (3.109)
2 2 3
For SU (4),

1 1 1 √ 1 1√ 1
I3 = λ3 , B = λ0 , C = (λ0 − 6λ15 ), S = √ λ8 − ( 6)λ15 − λ0
3 3 4 3 2 4

1 1√ 1 1
Q= λ0 + ( 3)λ8 − √ λ15 + λ0 (3.110)
2 2 6 6
Unitary Symmetry 97

Kindly observe the basic difference in the structure of the above electric
charge in SU (4) with respect to the corresponding SU (3) definition. It has
λ0 (the 4×4 unit matrix) which is not present in the SU (3) expression. These
are phenomenological fits of SU (4) to the known charges. Note also that in
the above we have not provided any expression for Y , the SU (3) hypercharge.
This has been a source of confusion in particle physics. Group theory is of no
help and everyone chooses what they like. For example, for SU (4) [[32] (page
280)] defines Y = B + S. Hence for them,

(B + S + C) (Y + C) 1 1√ 1
Q = T3 + = T3 + , Y = √ λ8 − ( 6)λ15 + λ0
2 2 3 2 12
(3.111)
Meanwhile [[33] (page 195)] defines (Y = S + B − C) with,

(B + S + C) (Y ) 1 1 1
Q = T3 + = T3 + + C, Y = √ λ8 + √ λ15 − λ0 (3.112)
2 2 3 6 6
Still other scientists use Y = 0 for the c-quark. The situation is quite
messy. However, what is phenomenologically correct is the expression in Equa-
tion 3.107. There the C-number is related to λ0 , and λ15 , the third diagonal
generator in SU (4). Similarly in SU (5) the beauty-number would be related
to λ0 and λ24 (the new diagonal generator in SU (5)) and similarly to λ0 and
λ35 in SU (6). These just add in the integral beauty and top quantum numbers
to obtain the correct charges as a generalization of the Gell-Mann-Nishijima
exxpression of the quark charges as
B+S+C +b+t
Q = I3 + (3.113)
2
Hence group theory is consistent with the above empirically verified defini-
tion of all the 6-flavoured (u, d, s, c, b, t)-quarks,. But it does not prove these.
Actually all the charges are correctly predicted only in the Standard Model
based on the group SU (3)3 ⊗SU (2)2 ⊗ U (1) [34], [35]. The above SU (n) flavor
definition is merely consistent with it.
What group theory predicts is that the SU (n) charges be defined by trace-
less matrices (Aij ) for that group. These evaluate correctly for SU (3) only,
mismatching the quark charges in SU (4)F . So the SU (3) success must be
accidental (or for some deep reason which we are as of yet unable to under-
stand) and these charges are clearly not the quark charges. The question one
may raise are these group theoretical charges completely irrelevant to nature?
Obviously they do not describe the 4-flavour quark charges. But what about
some other physical reality where these numbers may prove useful?
We are well aware that the same mathematical Lie group works simulata-
neously to describe different realities in nature. For example, the mathematical
SU (2) group maps the spin degrees of freedom, the isospin degrees of free-
dom and the nusospin degrees of freedom quite well [36]. Note all these are
98 Group Theory in Particle, Nuclear, and Hadron Physics

different and independent aspects of the physical reality. Also the mathemat-
ical framwork SU (3) works well for three flavours and also for three colours
in QCD. So could it be that these group theoretical SU (n) charges may be
found suitable in describing some other physical reality?
Note that here the electric charges are defined in terms of the two diagonal
generators of the group SU (3) . We have already seen that the electric charges
defined in terms of only the diagonal generators do not work for four flavours.
Hence let us abandon using these charges for the quarks. So let us assume that
there exist hypothetical new entities – let us call them h-electrons – which
do correspond to these new charges described in terms of A11 generators of the
SU (n) group. Here we shall define them in terms of the diagonal generators
of the SU (n) group. This method which was already used in 1983 [37], has
more predictive power.
So what is the most general intrinsic group theoretical definition of the
electric charge for the group SU (N ) for an arbitrary number N ?
Let us first re-write the Gell-Mann-Nishijima definition of the electric
charge for the group SU (3) for h-electrons as

Y3 (h3 + S3 )
Q3 = T3 + = T3 + (3.114)
2 2
where Q3 =Q, Y3 =Y, h3 =B and S3 =S are as per the standard notation. The
subscript 3 is used to indicate that it holds for the group SU (3). Here the
electric charge is defined in terms of the two diagonal generators of the group
SU (3). We take this as a fundamental property which defines the electric
charge consistently within the particular group under consideration, and uti-
lize it to define electric charges for any arbitrary value N for the group SU (N ).
It is that we should be able to define electric charge in terms of the similarly
defined (N − 1) diagonal generators of the group SU (N ). So for the group
SU (4) define:
Y3 Y4
Q4 = T3 + + (3.115)
2 3
where Y4 is to SU (4) what Y3 (or hypercharge) is to SU (3). Its values for the
h-leptons are 1/4, 1/4, 1/4, -3/4 (see Table 3.1).
For example, note that Y3 = h3 + S3 where h3 (=B) is the SU(3) h-lepton
number and S3 (=S) is a new quantum number which distinguishes it from the
other two h-electrons, h1 and h2. Similarly in SU (4)F let us define Y4 = h4 +S4
where we identify h4 with the SU (4) h-electron number and S4 with a new
quantum number which will distinguish the fourth h-electron from the others.
So Y4 = h4 + S4 looks like this:
     
1/4 1/4 0

 1/4  
= 1/4  
+ 0 

 1/4   1/4   0 
−3/4 1/4 −1
(3.116)
Unitary Symmetry 99

That is, the h-electron number h4 is 1/4 for all the h-electrons h1,h2,h3,h4.
The electric charge Q4 for the same h-electrons is 3/4, -1/4,-1/4,-1/4, respec-
tively. These are displayed in Table 3.1.
We extend this approach to SU (5) right away:
Y3 Y4 Y5
Q5 = T3 + + + (3.117)
2 3 4
Here Y5 is to SU (5) what Y4 and Y3 are to SU (4) and SU (3), respec-
tively. Also Y5 = h5 + S5 where h5 is the h-electron number and S5 is the
quantum number which distinguishes the h5 h-electron from all the others.
Now the charges for h1,h2,h3,h4,h5 in SU (5) are 4/5, -1/5, -1/5, -1/5, -1/5,
respectively.
These arguments are generalised to any SU (N ) for arbitrary h-electrons.
Clearly therein the charge QN of the h-electrons would be (NN−1) and - N1 .
Evidently same holds for two h-electrons also (see Table 3.1).
Now let us assume that the above SU (N ) is a valid model for finite and
arbitrary N as well as for h-electrons. Then these h-electrons that we obtain
here are having the h-electron number N1 and the electric charge - N1 . Hence
for example for N =5, the h-electron has a lepton number 1/5 and the electric
charges 4/5 and -1/5.
The next question pertains to whether this unique prediction of the electrc
charges for the Lie groups SU (N ) is of academic interest only, or if there
in fact exists experimental evidence of these (NN−1) and - N 1−1 charges. Most
intersestingly, indeed there is!
Completely unexpected effects in the transport properties of two-
dimensional electron gases when subjected to low temperatures and strong
magnetic fields were detected in the so-called Fractional Quantum Hall Effect
(FQHE) and Integral Quantum Hall Effect. These brought forth issues related
to macroscopic and microscopic quantum mechanics. All kinds of fractional
charges with odd denominators like, 1/2, 2/3, 3/5, 4/5, 2/7, 3/11, 6/23, 2/9,
10/21 plus many others with even-denominators like 5/2, have been experi-
mentally identified.
Earlier one knew of fractional quark charges of magnitude 2/3 and -1/3.
Now we are being confronted with all kinds of new fractional charges in FQHE.
This is certainly one of the most puzzling issues in physics today. It is pop-
ularly felt that FQHE implies the existence of quasi-particle collective states
within the broad framework of macroscopic quantum mechanics. For exam-
ple in Laughlin’s model [38], FQHE arises within a Gedanken experiment
approach to the gauge transformation of wave functions in a magnetic field.
His approach emphasizes a kind of macroscopic quantum phenomenon which
however says nothing about the actual quantization process.
We’d like to refer to Laughlin’s quotation [39]. “And I see the fractional
quantum Hall effect as a deep and important procedure for our guidance.
Its fractionally charged excitations are, I believe, related to the fractionally
charged quarks of the Standard Model of particle physics.”
100 Group Theory in Particle, Nuclear, and Hadron Physics
TABLE 3.1: Fractional charges of h-electrons in the groups
SU (2)h2 to SU (4)h4 as in FQHE
I2 h2 S2 Q2 Y3 h3 S3 Q3 Y4 h4 S4 Q4
h1 1/2 1/2 0 1/2 1/3 1/3 0 2/3 1/4 1/4 0 3/4
h2 -1/2 1/2 -1 -1/2 1/3 1/3 0 -1/3 1/4 1/4 0 -1/4
h3 0 0 0 0 -2/3 1/3 -1 -1/3 1/4 1/4 0 -1/4
h4 0 0 0 0 0 0 0 0 -3/4 1/4 -1 -1/4
h5 0 0 0 0 0 0 0 0 0 0 0 0

Indeed, here we make this conjecture into a concrete and specific math-
ematical structure which actually relates the quark fractional charges to the
fractional charges of the FQHE. An additional characteristic of our model
here is that it accounts for all the fractional charges already identified ex-
perimentally and makes predictions for many more to be discovered in the
future. Moreover, it also explains the anomalous fractional charge of 5/2 in a
consistent manner. It also unifies the FQHE with the IQHE as well.
We therefore have a unique correspondence with the charges of the FQHE.
Hence the FQHE of charge 4/5 (as quoted in the literature) would require to
put 4 h-leptons together as a composite whole. Also the FQHE of 5/2 is a
simple composite of five charged 1/2 h-electrons. Note that this picture is
general enough to account for any observed fractional charges. And of course
this model predicts many more to be discovered in the future. However, frame-
works incorporating the above success have to be implemented.
One point to be emphasised here is that as of now, there exists no single
model which accounts for all the experimentally determined FQHE charges.
Acceptable predictions of new fractional charges also do not exist. However,
the model presented here is unique in the sense that it is the only model which
accounts for all known fractional charges and also makes definite preditions
for other fractional charges to be discovered.

3.7 Appendix C: Linear Vector Space


A vector space is a set L of elements {x1 , x2 , , , , , }, called values of the
set L. This is a linear vector space if any element can be multiplied by a
complex number β and can be added together while the following conditions
hold [23], [20], [26].

a. If x1 , x2 ∈ L, then x1 + x2 ∈ L
b. (α + β)x = αx + βx
c. (αβ)x = α(βx)
Unitary Symmetry 101

d. 1 · x = x
e. α(x1 + x2 ) = αx1 + αx2
f. T here is a null vector (zero elements) 0 in L,
such that x + 0 = x f or all x ∈ L (3.118)

A set of vectors is linearly dependent if any member of the set may be


written as a sum of the other members of the set. If this is not possible then
the set is denoted linearly independent.
Let ( e1 , e2 , .....en ) be a set of n-linearly independent vectors in an n-
dimensional vector space. Thus, an arbitrary rector ~x can be written as a
linear combination of ~ei :
n
X
~x = xi~ei (3.119)
i=1

Here ~ei are termed base vectors and are said to form a basis or coordinate
system and xi are the components of ~x in this system.
In quantum mechanics the ~ei basis could be the state vectors |φ1 >
, ....|φn > which form a complete set of states in Ln . Thus any state vector
|φ > can be expressed as:
X
|φ >= ai |φi > (3.120)
i

Now we transform the coordinate system to another orthogonal form


(e01 , e02 ) then,
X
~x = x1 0 ~ei 0 (3.121)
i

Let us have a coordinate system in two dimensions and a vector ~x in it.


Let us perform a passive transformation that is a rotation of the entire
coordinate system while the vector itself is untouched. (The equivalent and
opposite rotation transformation is that of the vector itself and is called the
active transformation).
We can show that (see Figure 3.6)

x01 = cosθ x1 + sinθ x2


(3.122)
x02 = −sinθ x1 + cosθ x2

One can demononstrate that for the base e1 , e2 ,

e1 = cosθ e01 − sinθ e02


(3.123)
e2 = sinθ e01 + cosθ e02
102 Group Theory in Particle, Nuclear, and Hadron Physics

C G
2 6
2 2

X
2

6
1
x
e X

X
X l
2

0
0 0
x
i e
i Si

(a) (b)

FIGURE 3.6: Passive transformation on a two-dimensional vector

Verifying that x = xi ei = x0i e0i (here summation over the repeated index
is understood),
x1 cosθ e0i − x1 sinθ e02 + x2 sinθ e01 + x2 cosθ e02
= x1 e1 + x02 e02
0 0

The above equations  can be expressed


 in a matrix notation:
cosθ −sinθ
(x01 , x02 ) = (x1 , x2 )
sinθ cosθ
    0 
e1 cosθ −sinθ e1
and =
e2 sinθ cosθ e02
   
a11 a12 cosθ −sinθ
if a= =
a21 a22 sinθ cosθ
x01 = aji xj (3.124)
ei = aij e0j (3.125)
These equations are equivalent. Thus
x = xj ej = xj aji e0i = x0i e0i
Note that the above can also describe the case where a vector is trans-
formed to a different vector, i.e., as an active tranformation. Only that θ → −θ
and so xi = aij xj . This is a linear transformation of a vector x to another
vector x0 .

A : V → V, x0 = A(r) (3.126)
where A is a matrix (aij ) and x ∈ V, x0 ∈ V

Note that the linear transformation satisfies


Unitary Symmetry 103

A(ax + by) = aA(x) + bA(y) (3.127)


Thus the above transformation is linear.
To represent this linear transformation as a matrix, note that we require
a basis (ei ) for V . For x ∈ V given as x = xi ei the linearity condition needs

A(x) = xi A(ei )

A(ei ) being a vector is V , we have

A(ei ) = aij ej (3.128)

where (aij ) is n × n matrix.


Next given x ∈ V transformed to x ∈ V as

x0 = A(x) = A(xi ei ) = xi A(ei ) = xi aij ej

= xj aji ei

so
x0i = aji xj

which is the same as Equation 3.6.

Let us generalize this discussion to an arbitrary n-dimensional linear vector


space V with an element e = vi ei , where the basis vector ei , (i = 1, 2, ....n).
This is termed a contravariant vector if the components vi transform as

vi0 = aji vj (3.129)


under a change of basis as

ei = aij e0j (3.130)

Note that in Equation 3.129 it is row vectors that are summed over while
in Equation 3.130 the summation is over the column numbers. Because of this
difference in summing, the transformation is denominated contravariant. A
vector is said to transform covariantly if it transforms in the same manner as
the basis vectors.
Note that the above assumes that the vector is present in an underlying
coordinate space (e.g., a force field F~ (~r) at a position ~r). Such vector fields
104 Group Theory in Particle, Nuclear, and Hadron Physics

are assigned a value at each point in space. This is an example of what is


termed a vector bundle. The underlying space is denominated a manifold.
For irreducible linear transformations, we obtain

vi0 = (a−1 )ji vj0 (3.131)

e0i = (a−1 )ij ej (3.132)


It is imperative to note that an opposite convention is also utilized in
the literature. One may begin with Equation 3.131 and 3.132 as the starting
transformation and Equation 3.129 and 3.130 used as inverse of the same.
This is often used in literature. Given

x = xi ei (3.133)
(In quantum mechanics e1 , .....en may be basis vectors |φ1 >, |φ2 >
, . . . |φn > with |φ >= ai |φi >)

Let us change to a new coordinate system e0i through means of the B =


(bij ) transformation:

e0i = bij uj (3.134)


If the vector ~x is fixed while coordinate ei changes

x = xi ei = x0i e0i
So ei = (B −1 )ij e0j

As e0i are linearly independent,

x0j = xi (B −1 )ij (3.135)


∼ −1
Take A = B , then x0i = aij xj
If ~e is column vector found from ei ,

~e0 = B~e

∼ −1
~x0 = A~x = B ~x (3.136)
∼ ∼ −1
For the orthogonal transformation B = B −1 whence A = B = B and
so covariant and contravariant vectors are identical.
Now that the matrix representation of A changes under a change of basis
ei = cij e0j . Then matrix a0 is
Unitary Symmetry 105

A(e0i ) = a0ij e0j (3.137)


Next,

A(e0i ) = A((c−1 )ik ek ) = (c−1 )ik A(ek ) = (c−1 )ik akl el = (c−1 )ik akl clj e0j
(3.138)

where ej = cij e0j used in the last step. This yields:

a0ij = (c−1 )ik akl clj (3.139)


Which in matrix notation is

a0 = c−1 ac (3.140)
0
This is termed a similarity transformation of a → a . So matrix a and
a0 are similar if they are related as in Equation 3.140 by an invertible matrix c.

Diagonalization of a square matrix n × n representing a linear operator


A acting on a vector ~x represented by a column matrix X of order m × 1. This
transformation is →

y = A→ −x = λ→−
x where λ is the eigenvalue.

Ax = λX
or (A − λI)X = 0
X (3.141)
or (aij − λδij )xj = 0
j=1

For the eigenvectors to be non-trivial we read

|A − λI| = 0 (3.142)
This determinant is called the characteristic equation for the linear
operator A. It is also denominated the secular equation. The eigenvalues
are the roots of this equation.
a11 − λ a12 . . a1m


a21 a22 − λ . . a2m

A − λI = . . . . . =0


. . . . .

am1 am2 . . amm − λ
= λm + bm−1 λm−1 + . . . b1 λ + b0
is the characterestic polynomial of A. This is written as

|A − λI| = k(λ − λ1 )(λ − λ2 ) . . . (λ − λm )


106 Group Theory in Particle, Nuclear, and Hadron Physics

For real or complex numbers λ1 , λ2 , . . . λm and k. Therefore a polynomial


of degree m may have m complex roots. These m eigenvectors yield the spec-
trum of A with distinct eignevalues. Some roots may be degenerate with a
multiplicity n,

(A − λI) = (λ − λ1 )n g(λ)
The eigenvectors corresponding to degenerate eigenvalues need not be de-
generate themselves.
Even if the eigenvalue is zero the corresponding eigenvector is not zero.
Also eigenvectors are not unique, as the scalar multiple of an eigenvector is
also an eigenvector with the same eigenvalue.
Since if Aφi = λφi , Aφj = λφj

Then A(αφi + βφj ) = λ(αφi + βφj )

Also if A→−
x = λ→−
x then An = λn x as well
and −1 →

A (A x ) = λ(A−1 →

x ) so A x = λ−1 →
−1 →
− −
x

Example: Consider the active transformation by an angle θ (in the two-


dimensional rotation case)

 

− cosθ −sinθ
x 0 = A→

x with A =
sinθ cosθ
let A→−x = λ→−
x
cosθ − λ −sinθ
So A − λI =
=0
sinθ cosθ − λ
(cosθ − λ)2 + sinθ2 = 0

with λ1 = cosθ + isinθ


 λ2 = cosθ −isinθ
 
cosθ − α −sinθ x1
For eigenvector =0
−sinθ cosθ − α x2
giving x1 (cosθ − λ) + x2 sinθ = 0
. −x1 sinθ + x2 (cosθ − λ) = 0
 
1
For λ1 : x2 = −ix1 = −iα and eigenvectors as α
  −i
1
For λ2 : x2 = ix1 = iβ and eigenvectors β
i
(note eigenvectors, belonging to complex eigenvalues, are complex).
For degenerate eigenvalues proper use of orthonormality of the eigenectors
is required, e.g., to determine the eigenvector and eigenvalues of a matrix
operator:
Unitary Symmetry 107
 
1 1 1
A= 1 1 1 
1 1 1
Solution: Take A→−
x = λ→

x so (A − λI)x = 0

1−λ 1 1

Secular equation is |A − λI| = 1 1−λ 1 =0

1 1 1−λ
2
giving λ (λ − 3) = 0

With three solutions λ = 3 and degenerate λ2 = 0, λ3 = 0 having algebraic


multiplicity of two.
We need to solve
(1 − λ)x1 + x2 + x3 = 0
x1 + (1 − λ)x2 + x3 = 0
x1 + x2 + (1 − λ)x3 = 0
For λ = 3, x1 = x2 = x3 and the general eigenvector solution is
 
1
a 1 
1
For λ = 0 we have x1 + x2 + x3 = 0 which provides an infinite number of
choices. Note that A is a real symmetric matrix, which has two independent
eigenvectors corresponding to λ = 0. Let us choose x1 = 0 with x2 = −x3 and
this eigenvector:
 
0
b 1 
−1
The  other eigenvector
 of λ = 0 is
1
y= c 
−1 − c
Orthogonality implies bc + b + bc = b(1 + 2c) → c = − 21
so it 
is,   
1 2
d  − 12  = d2  −1 
− 12 −1
(Note that the normalized eigenvectors are with a = √13 , b = √12 , d2 =
√1 ).
6
Below we show that similar matrices have the same eigenvalue. Note that
A0 and A are similar matrices:

A0 = C −1 AC
108 Group Theory in Particle, Nuclear, and Hadron Physics

Eigenvalues of matrices A0 arise from the characteristic equation:

|A0 − λI| = |C −1 AC − λC −1 IC| = |C −1 (A − λI)C|


= |C −1 ||A − λI||C|
= |C −1 ||C||A − λI| = |C −1 C||A − λI)
= |A − λI|

Thus the similarity transformation preserves the eigenvalues of A.


As we saw that the similar matrices have the same eigenvalues. Hence a
matrix A can be diagonalized by using similarity transformation. Then,

C −1 AC = D
When D is the new diagonal matrix.
The diagonal matirx C is constructed by putting the eigenvectors of the
matrix A as its columns. As C is irreversible only if its column vectors are
linearly independent, thus the columns of C consist of these n-linearly inde-
pendent eigenvectors. Such a matrix so called the modal matrix of A.

————————————————–
Problem C.1: Determine the matrix which diagonalises the following
matrix and
 give its diagonal 
form,
2 0 3

M = 0 2√ i 3 
3 −i 3 3
————————————————–

3.8 Solutions of Problems


Solutions 3.1:

T T
X 0 = R(θ)X, X 0 = X T R(θ)† , X 0 X 0 = X T XR(θ)† R(θ)
 x0
     
 x cos θ sin θ cos θ − sin θ
x0 y 0 = x y
y0 y − sin θ cos θ sin θ cos θ
2
cos2 +θ sin θ
 
02 02 2 2
 − cos θ sin θ + sin θ cos θ
x +y = x +y
− sin θ cos θ + cos θ sin θ sin2 θ + cos2 θ
= x2 + y 2 1


————————————————–
Unitary Symmetry 109

Solution 3.2:
(a) First note that

X Xj
ex = 1 +
j=1
j!

X (X j )∗ ∗ † †
(eX )† = 1 + ; now X j = X ∗j → X j = eX
j!
As U is unitary
† †
(eiH )† (eiH ) = e−iH eiH = ei(H −H)
=1
So H † = H is Hermitian. Thus we can always write U = eiH .
(b) First identify det eX = etrX . As X is diagonalizable matrix S −1 XS =
∧ So X = SλS −1 where λ is the diagonal matrix with eigenvalues λ1 , λ2 , ....λn .

−1 X (SλS −1 )j X Sλj S −1
exp X = eSλS = 1+ = S1S −1 + = S (exp λ) S −1
j
j! j!

det(exp x) = det S · det(exp λ) det S −1 = det(exp λ) = eλ1 +λ2 +···λn = eT rX

As det U = det eiH = eiT rH = 1= 0. So Tr(H).


(c): H = H † . So along the diagonal H is real and hence specified by n-
parameters. But as T r(H) = 0, of these only (n − 1) generators are diagonal.
————————————————–
Solution 3.3:
O(n) is represented by n × n matrices of real numbers. So n2 number of
elements. Since orthogonal R−1 = RT and RRT = RT R = 1 . This has the
n-diagonal element equal to 1. So n such conditions are left with n2 − n terms.
The upper and lower off diagonal elements are not independent. There are
2
thus (n 2−n) such conditions.
So the number of independent elements of O(n) are
 (n2 −n)
n2 − n − 2 = n(n−1)
2 . So O(n) has n(n−1)
2 independent elements.
−1 T
Note as R = R , det R = +1 is not an independent condition. Consequently
SO(n) has n(n−1)
2 elements as well.
————————————————–
Solution 3.4:
Expanding the LHS up to the third order term yields

i(aµ Xµ )3 i(bν Xν )3
  
1 1
1 + iaµ Xµ − (aµ Xµ )2 − + ··· 1 + ibν Xν − (bν Xν )2 − + ···
2 6 2 6
110 Group Theory in Particle, Nuclear, and Hadron Physics

1 1
= 1 + iaµ Xµ + ibν Xν − (aµ Xµ )2 − (bν Xν )2 − (aµ Xµ )(bν Xν ) + · · ·
2 2

i(aµ Xµ )3 i(bν Xν )3 i i
− − − (aµ Xµ )2 (bν Xν ) − (aµ Xµ )(bν Xν )2 + · · ·
6 6 2 2
Now expand the right-hand exponent and retain terms up to the third
order to obtain
1 1
1 + iaµ Xµ + ibν Xν − (aµ Xµ ) (bν Xν ) − (bν Xν ) (aµ Xµ )
2 2

i 2
− ((aµ Xµ ) (bν Xν ) − 2 (aµ Xµ ) (bν Xν ) (aµ Xµ ) − (bν Xν ) (aµ Xµ )
12

2 2
+ (bν Xν ) (aµ Xµ ) − 2 (bν Xν ) (aµ Xµ ) (bν Xν ) + (aµ Xµ ) (bν Xν ) )

 2
1 1 1
+ −iaµ Xµ − ibν Xν − (aµ Xµ ) (bν Xν ) + (bν Xν ) (aµ Xµ )
2 2 2

1 3
+ {−iaµ Xµ − ibν Xν }
6
The last but one term is

1 2 2 i 2
[− (aµ Xµ ) − (aµ Xµ ) (bν Xν ) − (bν Xν ) (aµ Xµ ) − (bν Xν ) + (aµ Xµ ) (bν Xν )
2 2

i 2 i 2 i 2
− (aν Xν ) (bµ Xµ ) − (bν Xν ) (aµ Xµ ) − (aµ Xµ ) (bν Xν ) ]
2 2 2
The last term is

i 3 2 2
− [(aµ Xµ ) + (aµ Xµ ) (bν Xν ) (aµ Xµ ) + (bν Xν ) (aµ Xµ ) + (bν Xν ) (aµ Xµ )
6

2 2 3
+ (aµ Xµ ) (bν Xν ) + (aµ Xµ ) (bν Xν ) + (bν Xν ) (aµ Xµ ) (bν Xν ) + (bν Xν ) ]

Finally
Unitary Symmetry 111

1 2 1 2
= 1 + iaµ Xµ + ibν Xν − (aµ Xµ ) (bν Xν ) − (aµ Xµ ) − (bν Xν )
2 2

i 3 i 3 i 2 i 2
− (aµ Xµ ) − (bν Xν ) − (aµ Xµ ) (bν Xν ) − (aµ Xµ ) (bν Xν ) + · · ·
6 6 2 2
These prove the Campbell-Baker-Hausdorff formula.
————————————————–
Solution 3.5:
   
0 −1 0 −1
iθ 1i  θ
1 0 1 0
eiθX = e =e

θ2 −1 0 θ3 0 θ4
    
0 −1 1
=1+θ + + + 1 − ···
1 0 2! 0 −1 3! −1 0 4!

θ2 θ4 θ3 θ5
    
0 −1
=1 1− + − ··· + θ− + − ···
2! 4! 1 0 3! 5!
     
1 0 0 −1 cos θ − sin θ
= cos θ + sin θ =
0 1 1 0 sin θ cos θ
————————————————–
Solution 3.6:
det R = +1. Take

   
cos θ 0 sin θ cos θ 0 sin θ
R2 =  0 1 0  ; Cof R2 =  0 1 0 ;
− sin θ 0 cos θ − sin θ 0 cos θ
 
cos θ 0 − sin θ
Adj (Cof R2 ) =  0 1 0 
sin θ 0 cos θ
 
cos θ 0 − sin θ
R−1 =  0 1 0  = RT
sin θ 0 cos θ
Whence orthogonal.
————————————————–
112 Group Theory in Particle, Nuclear, and Hadron Physics

Solution 3.7:
[F1 , F2 ] = if123 F3
     
0 1 0 0 −i 0 0 −i 0 0 1 0
1 
4
1 0 0  i 0 0 − 14  i 0 0 1 0 0
0 0 0 0 0 0 0 0 0 0 0 0
     
i 0 0 −i 0 0 1 0 0
= 41 0 −i 0 − 14  0 i 0 = i 12 0 −1 0 = if123 F3
0 0 0 0 0 0 0 0 0
And so f123 = 1.
One finds the others as √
1 3
f147 = f246 = f257 = f345 = f516 = f637 = 2 f458 = f678 = 2
————————————————–
Solution 3.8:
By a trial and
q error method, we guess
2
λ(n2 −1) → n(n−1)
q q
For SU (2) it is +1, SU (3) → 13 , and so for SU (4) it is 16 and SU (5)
q
1
it is 10 .
The matrix itself is written by adding one more row and coloumn added
th
with ‘1’ placed along the first (n − 1)(n − 1) diagonal and −(n − 1) placed
th
along the (n)(n) diagonal to render it traceless.
————————————————–
Solution 3.9:
In SU (2), [J1 J2 ] = iijk Jk (i, j = 1, 2, 3) where ijk is the Levi-Civita
Tensor and is a completely antisymmetric entity, just like fλµν is in SU (3).
This fortuitous similarity is true because the indices in SU (2) run from 1 to 3.
Does it have any deep physical significance? It turns out that for SU (n) n > 3
the above completely antisymmetric property of the structure constants does
not survive. Basically this has to do with the diffference in the class structures
of the permutation symmetry S3 for SU (2) and SU (3) and say permutation
symmetry S4 for SU (4).
————————————————–
Solution 3.10:
 
1 0 0
I 2 = I12 + I22 + I32 = 2 0 1 0 = 1(1 + 1)1
0 0 1
Note I~2 is already diagonal:
    
1 0 0 π1 π1
I 2 |π >= 2 0 1 0 π2  = 2 π2 
0 0 1 π3 π3
Unitary Symmetry 113

    
0 −1 0 π1 −π2
I3 |π >= i 1 0 0 π2  = i  π1 
0 0 0 π3 0
We want a unitary transformation, |π 0 >= U |π > or
 0   
π1 U11 U12 U13 π1
π20  = U21 U22 U23  π2 
π30 U31 U32 U33 π3
Such that

I 2 |π 0 >= 2|π 0 >

I3 |π 0 >= a|π 0 >


Note the first one is always fulfilled.

I 2 U |π >= 2U |π >

U −1 I 2 U |π >= 2|π >


and LHS gives 2|π > too. For the second equation
  0  0
0 −1 0 π1 π1
i 1 0 0 π20  = a π20 
0 0 0 π30 π30
The secular equation is

−a −i 0

i −µ 0 = 0

0 0 −µ
whose solutions are a0 = 0, a+ = +1, a− = −1
Eigenvector for a = +1:
  0
−1 −i 0 π1
 i −1 0  π20  = 0
0 0 −1 π30
After normalization, up to a phase factor it is
 0  
π1 1
π20  = − √1  i 
π30 + 2 0

Similarly one finds,


114 Group Theory in Particle, Nuclear, and Hadron Physics

π10
     0  
1 π1 0
1
π20  = √ −i , π20  = 0
π30 − 2 0 π30 0 1
Hence the transformation from the Cartesian components to spherical com-
0
ponents yields (π+ , π00 , π−
0
)

0
√ √
π+ = −(π1 + iπ2 )/ 2 π00 = π3 π−
0
= (π1 − iπ2 )/ 2

  √1  
0 − 2 − √i2 0

π+ π1
 π00  = 
 0 0 1 π2 

0 √1
π− 2
− √i2 0 π3

Next we diagonalize I3 . Evaluate. < πµ |I30 | πν >= 𵆠I30 πν (µ, ν = +, 0, −)


P3
So πµ = n=1 Uµn πn where U is given above.
P3
So < πµ |I30 | πν >= n,m=1

Uµm πn† I30 πn Uνm
0
(3)
Inm = πn† I30 πm are as per the Cartesian representation above.
P3 0(3)
So < πµ |I30 | πν >= n,m=1

Uµn Inm Uνm

Introduce V = U ∗
P3 0(3) P3 0(3)
< π3 |I30 | πν >= n,m=1

Vµn Inm Vνm = n,m=1

Vµn Inm Vmν

or (I30 )diagnal−spherical = V I3,Cartesian


0
V†

Explicitly.
    √1  
− √12 √i 0 − 2 0 √i

2 0 −1 0 2 1 0 0
= i 0 0 1 1 0 0 − √i2 0 − √i2  = 0 0 0
   
√1 √i 0 0 0 0 0 1 0 0 0 −1
2 2

This unitary transformation V I3 V † diagonalizes the 3-component for the


spherical basis.
————————————————–
Solution C.1:
Take M →−
x = λ→ −
x . Note that it is a Hermitian matrix M = M † . So its
eigenvalues and eigenvector are real. Characteristic equation is
Unitary Symmetry 115


2−λ 0 3



|A − λI| = 0 2 −√λ i 3 =0

3 −i 3 3−λ

i.e., (λ + 1)(2 − λ)(λ − 6) = 0

So eigenvalues are λ1 = −1, λ2 = 2, λ3 = 6

Eigenvectors

−(2 − λ)x1 + 3x3 = 0



(2 − λ)x2 + i 3x3 = 0

3x1 − i 3x2 + (3 − λ)x3 = 0

Thus the normalized eigenvectors for the eiqenvalues -1,2,6 are, respec-
tively,
1 √ √ 1 √ 1 √
( 3, i, − 3), (i, 3, 0), √ (3, i 3, 4)
7 2 2 7
Hence the modal matrix C that diagonalizes M is

 q 
3 i 3

7 √2 2√ 7
 3 i √3

C= √i 
 q7 2 2 7 
3 4
− 7 0 √
2 7
 q q  q 
3 −i 3 i 3
7
√ − 37 7 √2

2√ 7
√7
C †C = 
  3 3 i √3

3
 − 2i 2√ 0   q7
  √
2 2 7


3 i √3 4

2 7
−2 7 √
2 7 − 37 0 4

2 7

The diagonal form


 of the matrix
 Mis 
−1 0 0 λ 0 0
MD = C † AC =  0 2 0  =  0 λ2 0 
0 0 0 0 0 λ3
————————————————–
Chapter 4
Permutation Symmetry

4.1 Indistinguishable Particles in Quantum Mechanics . . . . . . . . . . . . . . 117


4.2 Transpositions and Permutations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
4.3 Disjoint Cycles and Signatures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
4.4 Cayley’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
4.5 Symmetric Group S3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
4.6 Classes and Partitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
4.7 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132

4.1 Indistinguishable Particles in Quantum Mechanics


It is known that all electrons in the universe are identical to one another.
Two particles are said to be identical if all their intrinsic properties, e.g., mass,
electrical charge, spin, colour (as defined in QCD, the theory of the strong
interaction) are exactly the same. Similarly, all protons are identical to one
another, and so are all quarks. They possess the same defining properties under
similar interactions with each other. So far, the above definition can be used
to define identical entities both classically and quantum mechanically. Hence
we do talk of say, identical twins or identical billiard balls within a classical
setup. However, we can differentiate between two such identical entities by
placing labels (in principle at least) on each to differentiate between them.
Hence a black smudge mark affixed to one of the twins, or careful observation
of the trajectories of the billiard balls, would serve to distinguish these two
entities.
But in quantum mechanics we have to abandon this classical concept.
We cannot exactly follow individual particles on account of the uncertainty
principle. The best information we can obtain about some individual particle
location in space is that it exhibits a certain probability of being present there
at any given moment of time.
Let us now imagine two elementary particles, sufficiently well-separated
in space, so that there exists no overlap between their respective wave func-
tions. That is, they possess well-defined and well-separated probabilities of
being at their respective positions in space. Hence these two have distinct and
non-overlapping probabilities. So we can study their dynamics separately in

117
118 Group Theory in Particle, Nuclear, and Hadron Physics

quantum mechanics. Hence in this situation, we treat these two particles as


distinguishable, both quantum mechanically and of course classically [40].
Now let these two particles come closer, so that at some stage, the two wave
functions start overlapping in a significant manner. Quantum mechanically, it
is now no longer possible to distinguish between the two particles any more. It
is hence said that these two identical particles have become indistinguishable.
Note that this property of identical particles becoming indistinguishable is
a purely quantum mechanical reality. To belabour the point, the classical
concept of indistinguishability is non-existent.
Now we know that the proton is constituted of three quarks, its struc-
ture being (uud), namely one d- and two u-quarks. As all these are spin-1/2
fermioms we have to construct proper anti-symmetric functions from these
identical fermions. If you now point out that the two u- and d- type of quarks
are different from each other, then you deserve to be complemented. Well the
reason why these two may be considered as identical shall be explained in
Chapter 7.
Indistinguishability means that we cannot specify the exact location of
these particles. Thus it should be possible to permute their locations, and as
we exchange any two particles in a multi-particle state, we should not be able
to distinguish one from the other. As shall be elucidated below, the exchange in
position of identical particles has a so-called permutation symmetry. Clearly,
this permutation symmetry is extremely important for quantum mechanical
calculations of the multi-indistinguishable particles. Empirically it has been
determined that there are only two kinds of fundamental indistinguishable
identical particles: first, those whose intrinsic spin is half-integral like 1/2,
3/2, etc. denominated fermions; and second, those which exhibit integral spins
like 0,1,2, etc., which are termed bosons. Fermions have wave functions which
are antisymmetric under exchange of any two state labels and bosons are
symmetric under the same exchanges.
In an atom there may exist a large number of electrons which are identical
to each other. In a nucleus there may be Z number of protons and N number
of neutrons which are separately indistinguishable with respect to each other.
Not only that, including the concept of isospin in a nucleus, the neutron
and proton lose their individual separate Fermi character and they become
identical to one other. So knowing how a collection of fermions (or bosons)
behave is kinematically determined by their permutation symmetry.

4.2 Transpositions and Permutations


Let us start with some definitions.
Permutation Symmetry 119

Transposition: If G is any non-empty set, then the mapping of : G →


G is denominated a transposition. For example, for the set {1, 2} the possible
transpositions are

1 → 1 1 → 1 1 & 1 1 & 1
(a) ; (b) ; (c) ; (d) (4.1)
2 % 2 2 → 2 2 % 2 2 → 2

So a transposition is a mapping of the set onto itself. Next


Permutation: If G be a non-empty set, then a one-to-one mapping P: G
→ G is defined as a permutation.
For example, in the above example of the set {1, 2}, the transpositions (b)
and (c) are permutations.
The degree of a permutation P: G → G is the number of elements in the
finite (non-empty) set G. So in the above example, the degree of permutation
is two.
Let us consider the permutation of a set (1,2,3,4,5)
1→4,2→2,3→5,4→3,5→1
Let us now denote this permutation by
 
1 2 3 4 5
σ=
4 2 5 3 1
In this two row-line notation, the first row lists the elements of the set in
an un-permuted order in a sequential form. The second row lists the elements
in a permuted order, where the permutation is visible in each column. So
in the above example it is the permutation 1 → 4, 2 → 2, 3 → 5, etc. One
advantage of this two-row definition of permutations is that we can always list
the first row in any new arbitrary order, provided the same element as given
in the original permutation is listed in each column below it. Hence the above
permutation can also be written as
 
3 5 4 2 1
σ=
5 1 3 2 4
Next, let us take another independent permutation on the same set as
 
1 2 3 4 5
τ=
3 5 4 2 1
Take the product of these two permutations specified as τ σ. The convention
that we adopt here is that in this product, the term on the right is taken as
the first operation on the unpermuted set (1,2,3,4,5). On this set permuted by
the σ operation, we now act with the second operation τ to take the elements
to their final destination as per these operations.
Mathematically this is implemented as follows:
  
1 2 3 4 5 1 2 3 4 5
τσ =
3 5 4 2 1 4 2 5 3 1
120 Group Theory in Particle, Nuclear, and Hadron Physics
  
4 2 5 3 1 1 2 3 4 5
= − − − − −  − − − − − 
2 5 1 4 3 4 2 5 3 1
 
1 2 3 4 5

2 5 1 4 3
What we have accomplished above is the following. The first permutation
σ was left untouched. We rewrote the second permutation τ so that its first
row showed the same sequence as the second row of the first operation σ.
Next we visualized an imaginary line, indicated as dashes here, between the
two rows in σ and τ . The ingenious trick was to exploit this imaginary line
to convert the first and the second rows in the permutations into a mathe-
matical ratio with its characteristic numerators and the denominators. Next,
in a rather cavalier manner, we cancelled the first denominator (namely, the
first permutation) with the second numerator on the left. What was obtained
thereby, as illustrated above, is the final result of the product. It is impor-
tant to note that any product of two permutations on a set yields another
permutation defined on the same set. Hence the operation of the product of
permutations is closed on the relevant set.
As to what has happened above, it is imperative to observe how (τ σ) acts
on the position ‘1’:
(τ σ)1 = τ (σ1) = (τ 4) = 2
So (τ σ) takes the element in position ‘1’ to position ‘4’ and so on.
Symmetric Group
Let A be a finite set {1, 2, · · · n}, then the set of all the permutations of A
form a group called the symmetric group Sn . It is also termed the permuta-
tion group [41],[33]. Note in the set {1, 2, · · · n}, a particular number can be
permuted to ‘n’ places, the next one can only be permuted to the (n-1) left
places, the next one can be permuted to the remaining (n-2) places and so
on. Hence the total number of permutations is n(n-1)(n-2)...2.1= n! Thus the
number of elements (or the degree) of the symmetric group Sn is n!.
Now we prove that the symmetric group (or the permutation group) Sn
actually forms a group. We see how the various group conditions for Sn are
satisfied below.
1. Closure: Two successive permutations as a product, as seen above, yield
another permutation over the same set. So the property of closure is satisfied.
2. Associativity: It can be shown to be satisfied.
3. Identity: Clearly the identity element is:
 
1 2 3 ··· n
(4.2)
1 2 3 ··· n
4. Inverse: The inverse of the element is given as:
   
1 2 3 ··· n i j k ··· l

i j k ··· l 1 2 3 ··· n
Permutation Symmetry 121

We see that clearly,

    
i j k ··· l 1 2 3 ··· n 1 2 3 4 5
=
1 2 3 ··· n i j k ··· l 1 2 3 4 5

————————————————–
Problem 4.1: Demonstrate that the property of associativity holds for
the symmetric group.
————————————————–
Problem 4.2:
Given  
1 2 3 4
1 3 4 2
and assuming that the inverse is given by
 
1 2 3 4
a b c d
determine a,b,c,d explicitly.
————————————————–
Cyclic Notation A cycle of length n in compact notation is written as
(a1 ,a2 ,·,an ) where each element maps onto the next one and so on, except the
last one which is mapped onto the first one.
In our example above,
 
1 2 3 4 5
σ=
4 2 5 3 1
Here permutations are 1 → 4 → 3 → 5 → 1 and 2 which does not change.
So we can break this up and σ is rewritten as:
  
1 4 3 5 2
σ= → (1435)(2)
4 3 5 1 2
This is denominated cyclic rotation. Very often an element with a single
cycle, like (2) above, is skipped in writing the cycles, and then it is assumed to
be present by implication. So e=(1)(2)(3)...(n) is the identity element. Hence
this way, the above may be written as σ = (1 4 3 5). The advantage of writing a
particular permutation as a cycle is clearly a more compact way of expressing
it. Meanwhile the two row-line notation as given above is more transparent in
denoting multiplications. Hence, in practice one may jump from one notation
to the other as per ones convenience.
Below we write it in either of the ways, as
 
1 4 3 5
(1435) =
4 3 5 1
122 Group Theory in Particle, Nuclear, and Hadron Physics

Another independent permutation is


 
1 2 3 4 5
(1, 3, 5, 4)(2) =
3 2 5 1 4
Note that in cyclic notation, only the cyclic sequence is important.
So, (1,3,5,4) = (4,1,3,5) = (5,4,1,3) = (3,5,4,1) all are equivalent to each
other.
As another example, we multiply the following cycles in the symmetric
group S6 :

  
1 2 3 4 5 6 1 2 3 4 5 6
(2, 1, 5)(1, 4, 5, 6) =
5 1 3 4 2 6 4 2 3 5 6 1
  
4 2 3 5 6 1 1 2 3 4 5 6
=
4 1 3 2 6 5 4 2 3 5 6 1
 
1 2 3 4 5 6

4 1 3 2 6 5
And this result in cyclic notation is written as (142)(56)(3).

————————————————–
Problem 4.3:
Multiply (1 4 5 6) (2 1 5) in the symmetric group S6 and express the result
in cyclic rotation. See how it is related to what was shown above as (2 1 5) (1
4 5 6). Hence do they commute?
————————————————–

4.3 Disjoint Cycles and Signatures


Note that the identity element is cyclic of length 1. If we have different
cycles in a permutation and there is no element common to any two different
cycles, then all these cycles are termed disjoint.
Now in the previous example and the problem above, the cycles were not
disjoint as in (2, 1, 5) (1, 4, 5, 6) elements 1 and 5 were common to the two
cycle. Consider next,
 
1 2 3 4 5
= (1, 5) (2, 4, 3)
5 4 2 3 1
This permutation is representable in terms of disjoint cycles as none of the
elements in the two cycles are common.
Quite clearly,
Permutation Symmetry 123

(1, 5) (2, 4, 3) = (2, 4, 3) (1, 5)


That is, disjoint cycles commute. However,
 
1 2 3 4 5 6
(2, 1, 5) (1, 4, 5, 6) = = (1, 4, 2) (3) (5, 6)
4 1 3 2 6 5
Hence the above permutation can be written as a product of disjoint cycles.
In fact any permutation of a finite set is the product of disjoint cycles. A
cycle of length two is called a transposition.
Thus,

(123) = (13) (12)


  
1 2 3 1 2 3
=
3 2 1 2 1 3
    
2 1 3 1 2 3 1 2 3 
= = = 1 2 3
2 3 1 2 1 3 2 3 1
In fact any permutation can be expressed as a product of transpositions
but not necessarily in a unique manner,
    
1 2 3 = 1 3 1 2 = 2 3 1 3
   
= 1 2 1 3 1 2 1 3
or

        
3 4 5 6 = 3 6 3 5 3 4 = 3 5 6 4 5 6 4 5 3 4

Note that (123) can be expressed as different products of transpositions


but each of these is of an even order, for example, of order 2 and 4 here.
Next 3 4 5 6 can also be expressed as a product of an odd number of
transpositions, i.e., of order 3 and 5 above. So each cyclic permutation can be
expressed as a product of either an even number of transpositions or an odd
number of transpositions.

————————————————–
Problem 4.4:      
Verify that 3 4 5 6 = 3 5 6 4 5 6 4 5 3 4
————————————————–
Problem 4.5:
Write the following permutations as products of disjoint cycles:

     
3 1 4 2 1 2 3 4 5 1 2 3 4 5 6 7
(a) (b) (c)
4 2 1 3 2 4 3 5 1 4 2 1 3 7 6 5
124 Group Theory in Particle, Nuclear, and Hadron Physics

————————————————–

One can see that


     
1 3 2 4 1 2 3 4 = 1 4 2 3 (4.3)
Note that permutations in general do not commute, and so the order in
the cyclic products is important. Hence,

        
1 2 4 3 1 3 2 4 = 2 3 1 4 6= 1 3 2 4 1 2 4 3

The inverse of a permutation given in terms of disjoint cycles is the same


cycle in the reverse order. So,
−1
(1, 2, 3 · · · n) = (n · · · 3, 2, 1) (4.4)
Since

(n, · · · , 3, 2, 1) (1, 2, 3, · · · n) = (1, 2, 3, · · · n) (n, · · · 3, 2, 1)


= (1) (2) (3) · · · (n) = e
Thus,
−1
[(41) (2) (3)] = [(14) (2) (3)]

−1
[(7341) (825)] = (1437) (528)

————————————————–
Problem 4.6:  
1 2 3 4 5 6 7 8
Give the cyclic notation of the permutation
6 1 4 8 5 7 2 3
What is the order of this permutation?
————————————————–

The centre C of a group G is defined as the subset of elements c which


commute with all the elements of the group: C = {a ∈ G|cg = gc for all g∈G}.
Then we show that C is an Abelian subgroup of G.
Given c1 , c2 ∈ C means c1 c2 g = c1 gc2 = gc1 c2 so C is closed un-
der multiplication and therefore it is a subgroup of G. Take g = c2 then
cg = gc → cc2 = c2 c and hence this subgroup is Abelian.
Permutation Symmetry 125

————————————————–
Problem 4.7:
Confirm that the following permutation is a cyclic group of order five:
 
1 2 3 4 5
α=
2 3 4 5 1
————————————————–

In general any cyclic permutation can be written conveniently in terms of


(n − 1) transpositions as

(1, 2, 3, · · · n) = (1n) (n, n − 1) · · · (13) (12) (4.5)


(1, 2, 3, · · · n) = (12) (23) · · · (n − 1, n) (4.6)
Note that given a permutation of a finite set, the number of transpositions
in it is always either even or odd. A permutation is even or odd according to
whether it can be expressed as the product of an even number of transpositions
or the product of an odd number of transpositions.
m
We define signature as (−1) , where m is the number of transpositions.
The signature is +1 for an even permutation and −1 for an odd permutation.
Hence if permutations α, β ∈ Sn then the signature of the product αβ is equal
to the

m n m+n
( Signature of α).( Signature of β) ∼ (−1) · (−1) = (−1) (4.7)

(where m and n are the number of transpositions of α and β, respectively).


One can demonstrate that if a permutation is expressed in terms of two
different transpositions, then both are either even or both are odd. So the
signature of a permutation defines an intrinsic property. Hence all the per-
mutations of a permutation group Sn would separate out in terms of whether
they are even or odd. The total number of permutations of Sn are n!. It turns
out that half of them n! 2 constitute even permutations and the other half
constitute odd permutations. Identity of Sn clearly belongs to the set of even
permutations of order n! 2 . Thus this forms a subgroup of Sn and is called
the alternating group An
Since An is of order 2 in Sn it is a normal subgroup. Moreover, it acts as
a kernel of an isomorphic mapping to give a group of order 2. It is the factor
Sn
group A n
of order 2. Also all groups of order 2 are isomorphic to the groups
C2 or Z2 .
126 Group Theory in Particle, Nuclear, and Hadron Physics

4.4 Cayley’s Theorem


For the theory of finite groups the permutation group has a special place.
This is provided by Cayley’ s theorem, which states that every finite group of
order m is isomorphic to a subgroup of Sn , the permutation group of n-words
(n number of elements).
Note that for a finite group G, the multiplication of all the elements
(g : .i = 1, 2, .....n) by a given element ‘a’ is equivalent to permuting them.
Hence,

{ag1 , ag2 , · · · agn } = {gp1 , gp2 , · · · gpn } (4.8)


so that the element a of the group leads to the permutation P(a) of Sn ,
 
1, 2, · · · n
a : P (a) = (4.9)
p1 , p2 , · · · pn
As every index in the second row occurs once and only once, it is a per-
mutation. Now there are n elements agi (i = 1, 2, ....n) which in every permu-
tation must be distinct, since agi = agj multiplying a−1 agi = a−1 agj yields
gi = gj . But this is not possible as all the elements [gi ] are distinct. This is
also invertible as P (a) does not arise from any other group element a0 . So
agi = a0 gj → agi gi−1 = a0 gi gi−1 . So a = a0 .
Hence the correspondence a ↔ P (a) is one-to-one and it respects the
group structure of the groups G and Sn . Now a1 ↔ P (a1 ) and a2 ↔ P (a2 )
and a1 a2 ↔ P (a1 )P (a2 ). Next the n-permutations P (a1 ) is a subset of the
total n! permutation of Sn and forms a subgroup of Sn . Thus the group is
isomorphic to the group Sn .
Thus for C3 , P (e) = (1) (2) (3); P 2π = (123)and P 4π
 
3 3 = (132). Hence
C3 is isomorphic to this subgroup A3 , the alternating subgroup of S3 .

4.5 Symmetric Group S3


We prefer to use the word ‘symmetric group’ for S3 rather than the term
‘permutation group’.
Now S3 consists of elements {1, 2, 3} and has 3! = 6 members of the group.
Let us denote the members of this group as follows
   
1 2 3 1 2 3
ρ0 = , µ1 =
1 2 3 1 3 2
Permutation Symmetry 127

TABLE 4.1: Composition table for the symmetric group S3 ; rows specify the
first operation on the right in a product and the columns the second one from
the right.
. ρ0 ρ1 ρ2 µ1 µ2 µ3
ρ0 ρ0 ρ1 ρ2 µ1 µ2 µ3
ρ1 ρ1 ρ2 ρ0 µ2 µ2 µ1
ρ2 ρ2 ρ0 ρ1 µ3 µ1 µ2
µ1 µ1 µ3 µ2 ρ0 ρ2 ρ1
µ2 µ2 µ1 µ3 ρ1 ρ0 ρ2
µ3 µ3 µ2 µ1 ρ2 ρ1 ρ0

   
1 2 3 1 2 3
ρ1 = , µ2 =
2 3 1 3 2 1
   
1 2 3 1 2 3
ρ2 = , µ3 = (4.10)
3 1 2 2 1 3
The reason for the above names shall become clear in a moment. Check
  
1 2 3 1 2 3
µ1 ρ1 =
1 3 2 2 3 1
  
2 3 1 1 2 3
=
3 2 1 2 3 1
 
1 2 3
µ1 ρ1 = = µ2 (4.11)
3 2 1
  
1 2 3 1 2 3
ρ2 µ3 =
3 1 2 2 1 3
  
2 1 3 1 2 3
=
1 3 2 2 1 3
 
1 2 3
ρ2 µ3 = = µ1 (4.12)
1 3 2
The complete composition table of S3 is provided in Table 4.1.
Now refer to the Table 2.3, the composition table of D3 representing
the symmetries of an equilateral triangle. It is exactly identical to the com-
position table of the symmetric group S3 . For S3 we assigned names as
{ρ0 , ρ1 , ρ2 , µ1 , µ2 , µ3 } to correspond exactly to the appropriate geometric op-
eration for D3 − {ρ0 , ρ1 , ρ2 } which corresponds to the rotation subgroup C3
of D3 and {µ1 , µ2 , µ3 } which corresponds to the reflections in D3 . Note that
the operations in D3 are geometric operations on equilateral triangles, while
permutations are algebraic operations of shifting of position labels of points
128 Group Theory in Particle, Nuclear, and Hadron Physics

or particles. Hence isomorphism of the groups D3 and S3 should have deep


physical implications for theories which require these.
Note that the subgroup, the alternating group A3 of S3 , is given by the
set {ρ0 , ρ1 , ρ2 } provided in the upper left hand box of Table 4.1. It possesses
6
2 = 3 number of members and is isomorphic to C3 , the rotation cyclic group,
which is also a subgroup of C3 . Consequently, A3 ' C3 .
So our judicious selection of names for different permutations of S3 , re-
lating them to those of D3 , was performed in order to emphasise the specific
isomorphic between them.
In cyclic notation,

ρ0 = (1) (2) (3) , ρ1 = (123) , ρ2 = (132) (4.13)


µ1 = (23) , µ2 = (13) , µ3 = (12) (4.14)
So the subgroup structure of S3 is identical to that of D3 , which is illus-
trated in Figure 2.2. The cosets are also already known for D3 . A3 is thus the
S3
normal subgroup of S3 and the factor group A 3
is given in Table 2.8, which
is isomorphic to Z2 and C2 (Table 2.9).
Also S3 has 3 disjoint classes [ρ0 ] , [ρ1 , ρ2 ], {µ1 , µ2 , µ3 } of orders 1,2,3,
respectively. Remember that all the members in a class possess the same
signature; i.e., either these are all even or are all odd.
The isomorphing between D3 and S3 may suggest that D4 , the dihedral
group of a square (see Figure 2.1) may be related to S4 . But be warned, even
S2 ' C2 only as two members present, and also remember that D4 has 8
elements while S4 has 4! = 24 elements. So S2 is geometrically related to
the cyclic group of rotation C2 , and D3 ' S3 . But geometrically D4 is not
isomorphic to S4 . What geometric figure is S4 isomorphic to?

————————————————–
Problem 4.8:
For S4 give all the 4! = 4.3.2 = 24 elements in the two row rotation of the
symmetric group S4 . How many cyclic structures do these elements present?
Give the complete list in cyclic notation as well.
————————————————–

We saw that S2 ∼= C2 and S3 ≡ D3 . So S2 relates to the geometry of a line


in two dimensions, and S3 to the geometry of an equilateral triangular figure
in three dimensions. And so S4 (not being isomorphic to D4 ) should pertain
to one higher dimension whence S4 should be related to the geometry of a
tetrahedron (with each side equal to the other, see Figure 4.1). Here each of
the four faces forms an equilateral triangle.
In S3 note that the equilateral triangle is actually drawn on a plane (i.e.,
a two dimensional surface). But S3 requires one to go out of the plane to
obtain permutations µ1 , µ2 , and µ3 , and thus this necessitates a three dimen-
sional space. So in S4 though the figure exists in three-dimensional space,
Permutation Symmetry 129

2
4

FIGURE 4.1: Regular tetrahedron symmetry as isomorphic to the symmet-


ric group S4

there should be permutations in S4 which would require one to move to four-


dimensional space. These would be necessitated to flip over the tetrahedron
in order to obtain all required permutations. Take the permutation
 
1 2 3 4
(4.15)
1 3 2 4
From Figure 4.1 one can discern that this permutation corresponds to a
reflection in the plane containing the line through the vertices numbered 1
and 4, and perpendicular to the line through vertices numbered 2 and 3.

————————————————–
Problem 4.9:
If H is a subgroup of G; H ⊂ G then g1 ∈ G, then the conjugacy elements
are either in H or in g1 H. By considering these two elements and given that
the order of [h] = 21 [g] then demonstrate that H must be a self-conjugate
subgroup of G.
————————————————–
Problem 4.10:
The permutation group S5 has 5! = 5.4.3.2 = 120 elements. Write down
all the elements separated as per this cyclic structure.
————————————————–
130 Group Theory in Particle, Nuclear, and Hadron Physics

4.6 Classes and Partitions


We saw that S3 possesses three disjoint classes [ρ0 ], [ρ1 , ρ2 ], [µ1 , µ2 , µ3 ], of
order 1,2,3, respectively. In general all the permutations in the same class
(conjugacy class) have identical disjoint cyclic structure. Moreover, each class
is fully characterized by the disjoint cycle structure alone.
Now each permutation can be written as a product of closed cycles (dis-
joint), without common elements. So in Sn let us assume that a j-cycle appears
kj times. Thus one has (kj ≥ 0),

k1 + 2k2 + · · · + nkn = n (4.16)


As a class is defined by its cyclic structure, each set of integers kj above
corresponds to a class of Sn .
We write the above as

λ1 + λ2 + · · · λn = n (4.17)
with

λ1 = k1 + k2 + · · · + kn
λ2 = k2 + · · · + kn
···
λn = kn (4.18)
and with the condition λ1 ≥ λ2 ≥ · · · ≥ λn ≥ 0.
The set λ = [λ1 , λ2 · · · λn ] is called a partition. We write

k1 = λ1 − λ2
k2 = λ2 − λ3
···
kn = λn
Therefore the set {k1 , k2 · · · kn } is related in a one-to-one manner to the set
[λ1 , λ2 · · · λn ]. Note that the number of classes are separated by the partition
λ. Thus the number of the partition of n in Sn is equal to the number of
classes of Sn . Hence the usefulness of partitioning the number n is to yield the
number of partitions. So for S3 ,

3=3+0+0
3=2+1+0
3=1+1+1
Permutation Symmetry 131
 
Or we write the partition of 3 as [3] , [21] ,[111]. One may write [111] = 13 .
So for S3 the classes are

[3] ∼ [ρ0 ] , [21] = [ρ1 , ρ2 ] , 13 = [µ1 , µ2 , µ3 ]


 
(4.19)
For S2 there exist two classes as per the partitions (n=2, 1+1),

[2] , [11] (4.20)


Next for S4 the partition of n = 4 is

4 = 4 = [4]
4 = 3 + 1 = [31]
4 = 2 + 2 = [22] = 22
 

4 = 2 + 1 + 1 = [211] = 212
 

4 = 1 + 1 + 1 = [1111] = [14 ]
Consequently S4 has five classes. In the problem-set below we shall demon-
strate that these classes are: one containing all unit cycles [1111], i.e., the
identity which is always a class by itself; one containing the single transpo-
sition [2111] which has six elements; one containing only a pair of elements
having pairs of transpositions [2+2] and which has three elements; one con-
taining only one three-cycle permutation [3 + 1] (it has 8 elements); and one
consisting of all 4-cycle permutations [4] (which has 6 elements).

————————————————–
Problem 4.11:
From all the 24 permutations of S4 , obtain the class structure and the
number of elements in each class, by partitioning n = 4.
————————————————–
Problem 4.12:
By partitioning n=5 of S5 obtain the number of classes in S5 . This is
related to the cyclic structure of the elements. So match this result of the
cyclic structure of S5 obtained in Problem 4.10. Hence determine the number
of elements in each class.
————————————————–

Also useful is the formula defining the number of elements D in a class of


Sn characterized by ni cycles of length li :
n!
D= nl (4.21)
Πi li i(nli !)
So for S3 :
132 Group Theory in Particle, Nuclear, and Hadron Physics

3!
[3] , D = =2
3 · (1!)
3!
[21] , D = =3
2 · (1!) .1 (1!)
3!
[111] , D = =1 (4.22)
3!

————————————————–
Problem 4.13: Give the number of elements of each class of S4 and S5
using the above formula.
————————————————–
Problem 4.14: What classes do the alternating groups A3 ⊂ S3 and
A4 ⊂ S4 consist of?
————————————————–

4.7 Solutions of Problems

Solution 4.1:
We have to demonstrate that for any symmetric group of degree n, the
expression (A . B) . C = A . (B . C) holds true.
Let us judiciously define the permutations A, B, and C as follows:
     
b1 b2 ··· bn c1 c2 ··· cn d1 d2 ··· dn
A= , B= ,C=
a1 a2 ··· an b1 b2 ··· bn c1 c2 ··· cn

    
b1 b2 ··· bn c1 cn c2 ···
d1 d2 · · · dn
(AB)C = [ ]
a1 a2 ··· an b1 bn b2 ···
c1 c2 · · · cn
    
c1 c2 .... cn d1 d2 .... dn d1 d2 .... dn
= =
a1 a2 .... an c1 c2 .... cn a1 a2 .... an
    
b1 b2 .... bn c1 c2 .... cn d1 d2 .... dn
A.(B.C) = [ ]
a1 a2 .... an b1 b2 .... bn a1 a2 .... an
    
b1 b2 .... bn d1 d2 .... dn d1 d2 .... dn
= =
a1 a2 .... an b1 b2 .... bn a1 a2 .... an
————————————————–
Permutation Symmetry 133

Solution 4.2:
 
1 2 3 4
Define as the inverse element.
a b c d
     
1 2 3 4 1 2
1 3 4 2 3 4
1 2 3 4
=
a b c d 1 3
a c d b 4 2
1 3 4 2
 
1 2 3 4
=
a c d b
 
1 2 3 4
And this should be equal to =
1 2 3 4
So we find that a=1, c=2, d=3, b=4. Hence the required inverse is
     
1 2 3 4 1 2 3 4 1 3 4 2
= =
a b c d 1 4 2 3 1 2 3 4
————————————————–
Solution 4.3:
  
1 2 3 4 5 6 1 2 3 4 5 6
4 2 3 5 6 1 5 1 3 4 2 6
  
5 1 3 4 2 6 1 2 3 4 5 6
=
6 4 3 5 2 1 5 1 3 4 2 6
 
1 2 3 4 5 6
=
6 4 3 5 2 1
Thus in the cyclic notation it is (16)(245)(3).
————————————————–
Solution 4.4:

    
1 2 3 4 5 6 1 2 3 4 5 6 1 2 3 4 5 6
[ ]
1 2 5 4 3 6 1 2 3 6 5 4 1 2 3 4 6 5
  
1 2 3 4 5 6 1 2 3 4 5 6
[ ]
1 2 3 5 4 6 1 2 4 3 5 6
    
1 2 3 4 5 6 1 2 3 4 5 6 1 2 3 4 5 6
=[ ]
1 2 5 6 3 4 1 2 3 4 6 5 1 2 5 3 4 6
  
1 2 3 4 5 6 1 2 3 4 5 6
=
1 2 5 6 4 3 1 2 5 3 4 6
  
1 2 5 3 4 6 1 2 3 4 5 6
=
1 2 4 5 6 3 1 2 5 3 4 6
134 Group Theory in Particle, Nuclear, and Hadron Physics
 
1 2 3 4 5 6 
= = 3 4 5 6
1 2 4 5 6 3
————————————————–
Solution 4.5:

      
(a) 1 2 3 4 (b) 1 2 4 5 3 (c) 1 4 3 5 7 2 6

————————————————–
Solution 4.6:
  
1 6 7 2 5 3 4 8 . Each cycle has order 4 and 3 and LCM of
4 and 3 is 12. Hence this permutation exhibits an order of 12.
————————————————–
Solution 4.7:

     
1 2 3 4 5 1 2 3 4 5 2 3 4 5 1 1 2 3 4 5
α2 = =
2 3 4 5 1 2 3 4 5 1 3 4 5 1 2 2 3 4 5 1
 
1 2 3 4 5
=
3 4 5 1 2
   
31 2 3 4 5 4 1 2 3 4 5
α = , α =
4 5 1 2 3 5 1 2 3 4
 
1 2 3 4 5
α5 = =I
1 2 3 4 5
Consequently it is a cyclic group of order 5.
————————————————–
Solution 4.8:

       
1
2 3 4 1 2 3 4 1 2 3 4 1 2 3 4
(a) : , , ,
1
2 3 4 2 1 3 4 3 2 1 4 4 2 3 1
     
1 2 3 4 1 2 3 4 1 2 3 4
, , ;
1 3 2 4 1 4 3 2 1 2 4 3
       
1 2 3 4 1 2 3 4 1 2 3 4 1 2 3 4
(b) : , , ,
2 3 1 4 3 1 2 4 2 4 3 1 4 1 3 2
       
1 2 3 4 1 2 3 4 1 2 3 4 1 2 3 4
, , , ;
3 2 4 1 4 2 1 3 1 3 4 2 1 4 2 3
     
1 2 3 4 1 2 3 4 1 2 3 4
(c) : , , ;
2 1 4 3 3 4 1 2 4 3 2 1
     
1 2 3 4 1 2 3 4 1 2 3 4
(d) : , ,
2 3 4 1 2 4 1 3 3 4 2 1
Permutation Symmetry 135
     
1 2 3 4 1 2 3 4 1 2 3 4
, ,
3 1 4 2 4 3 1 2 4 1 2 3
The cycle structure of the above is

{(1) (2) (3) (4)}


{(1, 2) (1, 3) (1, 4) (2, 3) (2, 4) (3, 4)}
{(12) (34) , (13) (24) , (14) (23)}
{(123) , (132) , (124) , (142) , (134) , (143) , (234) , (243)}
{(1234) , (1243) , (1324) , (1342) , (1423) (1432)}
————————————————–
Solution 4.9:
If the conjugate element g ∈ G is in H then H = gHg −1 as H ⊂ G. Now
take g1 h1 = g. Thus ghg −1 = g1 h2 g1−1 (h1 , h2 ∈ H). It may be in H. If this
is not so then as g = g1 h1 it must be in g1 H with g1 h2 g1−1 = g1 h3 (h3 ∈ H).
This however leads to g1 = h−1 3 h3 ∈ H but we know g1 ∈ / H. Thus g1 H is
another conjugacy class of G in addition to H.
————————————————–
Solution 4.10:
 
1 2 3 4 5
(a) : = (1) (2) (3) (4) (5) = () : 1 element
1 2 3 4 5
     
1 2 3 4 5 1 2 3 4 5 1 2 3 4 5
(b) : = (12) ; = (13) ; = (14) ;
2 1 3 4 5 3 2 1 4 5 4 2 3 1 5
     
1 2 3 4 5 1 2 3 4 5 1 2 3 4 5
= (15) ; = (23) ; = (24) ;
5 2 3 4 1 1 3 2 4 5 1 4 3 2 5
     
1 2 3 4 5 1 2 3 4 5 1 2 3 4 5
= (25) ; = (34) ; = (35) ;
1 5 3 4 2 1 2 4 3 5 1 2 5 4 3
 
1 2 3 4 5
= (45) ; 10 elements
1 2 3 5 4
   
1 2 3 4 5 1 2 3 4 5
(c); = (12) (34) ; = (12) (35) ;
2 1 4 3 5 2 1 5 4 3
   
1 2 3 4 5 1 2 3 4 5
= (12) (45) ; = (13) (24) ;
2 1 3 5 4 3 4 1 2 5
   
1 2 3 4 5 1 2 3 4 5
= (13) (25) ; = (13) (45) ;
3 5 1 4 2 3 2 1 5 4
   
1 2 3 4 5 1 2 3 4 5
= (14) (23) ; = (14) (25) ;
4 3 2 1 5 4 5 3 1 2
   
1 2 3 4 5 1 2 3 4 5
= (14) (35) ; = (15) (23) ;
4 2 5 1 3 5 3 2 4 1
136 Group Theory in Particle, Nuclear, and Hadron Physics
   
1 2 3 4 5 1 2 3 4 5
= (15) (24) ; = (15) (34) ;
5 4 3 2 1 5 2 4 3 1
   
1 2 3 4 5 1 2 3 4 5
= (24) (35) ; = (25) (34) ;
1 4 5 2 3 1 5 4 2 3
 
1 2 3 4 5
= (23) (45) ; 15 elements
1 3 2 5 4
(d) : (123) (45) ; (132) (45) ; (124) (35) ; (142) (35) ; (134) (25) ; (143) (25) ;
(125) (34) ; (152) (34) ; (135) (24) ; (153) (24) ; (145) (23) ; (154) (23) ;
(234) (15) ; (243) (15) ; (235) (14) (253) (14) ; (245) (13) ; (254) (13) ;
(345) (12) ; (354) (12) ; 20 elements
(e) : (123) ; (132) ; (124) ; (142) ; (134) ; (143) ; (125) ; (152) ; (135) ; (153) ; (145) ;
(154) ; (234) ; (243) ; (235) (253) ; (245) ; (254) (345) (354) ; 20 elements
(f ) : (1234) ; (1243) ; (1235) ; (1245) ; (1254) ; (1253) ; (1324) ; (1342) ; (1325) ;
(1345) ; (1354) ; (1352) ; (1423) ; (1432) ; (1425) ; (1435) ; (1453) ; (1452) ;
(1534) ; (1543) ; (1524) ; (1542) ; (1523) ; (1532) ; (2345) ; (2354) ; (2435) ;
(2453) ; (2534) ; (2543) ; 30 elements
(g) : (12345) ; (12435) ; (12354) ; (12453) ; (12543) ; (12534) ; (13245) ; (13425) ;
(13254) ; (13452) ; (13542) ; (13524) ; (14235) ; (14325) ; (14253) ; (14352) ;
(14532) ; (14523) ; (15342) ; (15432) ; (15243) ; (15423) ; (15234) ; (15324) ;
24 elements
————————————————–
Solution 4.11:
From Problem 4.8 on S4 , we have the complete cyclic structure and thus
the class structures are as follows:

[1111] : {(1) (2) (3) (4)} ; single member


[2 + 1 + 1] : {(1, 2) , (1, 3) , (1, 4) , (2, 3) , (2, 4) , (3, 4)} ; 6 members
[2 + 2] : {(12) (34) , (13) (24) , (14) (23)} ; 3 members
[3 + 1] : {(123) , (132) , (124) , (142) , (134) , (143) , (234) , (243)} 8mem.
[4] : {(1234) , (1243) , (1324) , (1342) , (1423) , (1432)} ; 6 members
Total : 1+6+3+8+6=24
————————————————–
Solution 4.12:
By partitioning n=5 there are the following 7 classes:
Permutation Symmetry 137

TABLE 4.2: Number of elements in classes of S4 and S5


S4 S5
[1111]; D = 144!(4!) = 1 [11111]; D = 155!(5!) = 1
4! 5!
[211]; D = 21 (1!)12 (2!) = 6 [2111]; D = 2(3!) = 10
[22]; D = 224!
(2!) = 3 [221]; D = 235!
(2!) = 15
1
4! 5
[31]; D = 31 (1!)1 1 (1!) = 8 [311] ; D = 3(2!) = 20
4! 5!
[4]; D = 4(1!) = 6 [32] ; D = 3.2 = 20
= 24 = 4! [4]; D = 5!
4 = 30
5!
[5]; D = 5 = 24
= 120 = 5!

5 = 5 = [5] , 24 members
5 = 4 + 1 = [41] , 30 members
5 = 3 + 1 + 1 = [311] , 20 members
5 = 3 + 2 = [32] , 20 members
5 = 2 + 2 + 1 = [221] , 15 members
5 = 2 + 1 + 1 + 1 = [2111] , 10 members
5 = 1 + 1 + 1 + 1 + 1 = [11111] , 1 members
Total: 1 + 10 + 15 + 20 + 20 + 30 + 24 = 120 = 5! of S5
————————————————–
Solution 4.13:
Solution in Table 4.2
————————————————–
Solution 4.14:
The group A3 consists of the three even permutations contained in the
class [111] = 13 and [3].
The group A3 consists of 12 even elements contained in the classes
[1111],[22], and [31].
————————————————–
Chapter 5
Young Diagram

5.1 Symmetries and Young Diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139


5.2 Young Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
5.3 Young Diagram and Unitary Symmetry . . . . . . . . . . . . . . . . . . . . . . . . . 151
5.4 Product of Irreducible Representations . . . . . . . . . . . . . . . . . . . . . . . . . . 160
5.5 Multiplets of the SU (n − 1) Subgroup of SU (n) . . . . . . . . . . . . . . . . 162
5.6 The Reduction SU (m + n) → SU (m) ⊗ SU (n) . . . . . . . . . . . . . . . . . 165
5.7 Reduction of SU (mn) → SU (m) ⊗ SU (n) . . . . . . . . . . . . . . . . . . . . . . 168
5.8 Coefficients of Fractional Parentage and Isoscalar Factors . . . . . . 171
5.9 Appendix D: Representation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
5.10 Appendix E: Symmetry in Quantum Mechanics . . . . . . . . . . . . . . . . 184
5.11 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189

5.1 Symmetries and Young Diagrams


In Section 4.1 we discussed the quantum mechanical problem of how to deal
with two or more than two identical particles. At the heart lies the issue of the
representation theory (Appendix D) as to what are the “good” states built up
of multi-identical particles and how to obtain their irreducible representations
(Appendix E).
Let us start with the simplest system of two spin entities (say electrons)
residing at points ‘1’ and ‘2’. Suppose their states overlap over a finite region
containing points 1 and 2. Then as each electron can have spin − up ↑ or
spin − down ↓ then the possible combinations of the two-electron quantum
mechanical wave function are,

↑ (1) ↑ (2)
↑ (1) ↓ (2)
↓ (1) ↑ (2)
↓ (1) ↓ (2) (5.1)
Now due to the quantum mechanical uncertainty principle we cannot say
whether it is the states ↑ or ↓ at either of the positions 1 or 2. So the above
states are not “good” quantum states as these states correspond to a reducible

139
140 Group Theory in Particle, Nuclear, and Hadron Physics

representation of the spin-group. Good quantum states are built up as linear


combinations of the above states (when we include proper normalization fac-
tors, too) or these states provide proper irreducible representations of the
spin-group. These are
1
↑ (1) ↑ (2), √ (↑ (1) ↓ (2)+ ↓ (1) ↑ (2)), ↓ (1) ↓ (2) (5.2)
2
1
√ (↑ (1) ↓ (2)− ↓ (1) ↑ (2)) (5.3)
2
Now on exchange of the spin states of the two electrons we obtain either
of the three states in Equation 5.2 which are either symmetric or they do
not change signs under exchange of particles; or they do change signs under
this exchange, as in Equation 5.3 and so are antisymmetric. As for ↑ (1) ↑
(2) SZ = +1, ↓ (1)+ ↓ (2) SZ = −1 and thus the middle one would
~ = 1. The one
correspond to SZ = 0 and therefore the three states represent S
~
in Equation 5.3 stands for the S = 0, SZ = 0 state. This reduction is given
in quantum mechanical text books as

~1 ~1
⊗ = ~1 + ~0 (5.4)
2 2
indicating that on joining two spin-halves we get two states of spin ~1 and ~0.
Now spin ~0 has two states – namely, it is two-dimensional, spin ~1 has three
states S2 = +1, 0, −1), namely, it is three-dimensional and spin 0 has only
one state, i.e., it is one-dimensional. Thus we express Equation 5.4 in terms
of dimensions as

2⊗2=3⊕1 (5.5)
This is a more appropriate representation of the above coupling in terms
of group theory.

− → − → −
Similarly three spin-halves yield the spin 23 , 21 , 12 as

2×2×2=4⊕2⊕2 (5.6)
And so on. Combining more than one spin in atoms and nuclei, combin-
ing three-quarks in a nucleon or a quark-antiquark in a meson or 6-quarks,
9-quarks and 12-quarks in A = 2, 3, 4 nuclear systems, respectively, are ba-
sic requirements in physics. For these we have to deal with the permutation
group Sn for different number of electrons, quarks, etc., and obtaining their
irreducible representations (i.e., good quantum mechanical states) is a long
and tedious task if n in Sn is large. A very powerful and practical as well as
labour saving graphical method, the so-called Young Diagram technique has
been found to be very useful.
Young Diagram 141

In the previous chapter we saw that for Sn the number of classes is speci-
fied by partitioning n as follows:

λ1 + λ2 + · · · + λh = n

λ1 ≥ λ2 ≥ λ3 · · · ≥ λh (5.7)
We denominate the partition [λ1 λ2 λ3 · · · λh ]
So for S2 and S3 , respectively,

S2 : n = 2 : [2], [11] (5.8)

S3 : n = 3 : [3], [21], [111] (5.9)


This partition gives the number of classes of Sn . So S2 has two classes and
S3 has three classes.
Now it turns out that this partitioning corresponds to the total number
of irreducible representations of the particle. So for two particles the total
number of irreducible representation is two, corresponding to the two classes
reduction of S2 .
Now each partition may be represented by a Young Diagram (YD). This
consists of arrangement of n-cells in h-rows. Each row begins with the same
vertical line on the left and the number of cells in successive rows are
λ1 λ2 λ3 · · · λh . Thus the Young Diagrams for S2 and S3 are

n = 2 : [2] ≡ ; [11] ≡ (5.10)

n = 3 : [3] ≡ ; [21] ≡ ; [111] ≡ (5.11)


Each partition defines a class of Sn . So the number of irreducible rep-
resentations of Sn is equal to the number of partitions or equivalently the
corresponding Young Diagrams. Thus S2 , S3 have 2 and 3 irreducible repre-
sentations, respectively.

————————————————–
Problem 5.1: Given the partition n = 4 of S4 find all the Young Diagrams
of it.
————————————————–

Consider S2 . Let the wave function of the two identical particles be given
142 Group Theory in Particle, Nuclear, and Hadron Physics

by φ(1, 2). The numbers 1 and 2 labelling particles are giving all the degrees of
freedom of the particles; say, these are the position, the spin, and the isospin.
Then, 1 ∼ (x1 , y1 , z1 , s1 , τ1 ) and 2 ∼ (x2 , y2 , z2 , s2 , τ2 ) (where s stands for spin
and τ for isospin). Now it may be possible to give φ(1, 2) as the product of
two identical particle states say φ(1, 2) = ψ(1)ρ(2) when ψ, ρ are orthogonal
identical particle states. However, it is not a prior conditional and φ(1, 2) may
not be separable (an example exists in the quark model where the octet wave
function is √12 (φρ χρ + φλ χλ ) where φ are SU (3)f wave functions and χ is the
spin SU (2) wave function ). However, due to quantum mechanical uncertainty
principle we cannot distinguish between φ(1, 2) and φ(2, 1). But we can build
a symmetric φS and an antisymmetric function φA as

φS = φ(1, 2) + φ(2, 1)

φA = φ(1, 2) − φ(2, 1) (5.12)


Thus 1 ↔ 2 the state φs does not change sign (and hence is symmetric
under this exchange) and φA changes sign (and thus is antisymmetric under
this exchange). Now we define permutation operator P12 as inducing transpo-
sitions,

φ(2, 1) = P12 φ(1, 2) (5.13)


Thus φS and φA are eigenstates of the permutation operator,

P12 φS = +φS
P12 φs = −φA (5.14)
So both the eigenstates φS and φA are non-degenerate or singlet of all
permutations of S2 . So φS and φA are separate basis states as one-dimensional
(singlet) representation of the permutation group.
Now take a single state as represented by a single box – as a Young
Diagram. Let us associate the two Young Diagrams of S2 as given in Equation
5.12 as

φS ≡ φA ≡ (5.15)
So boxes in a row represent a symmetric state and boxes in a column
represent an antisymmetric state.
Now above is a Young Diagram (YD). But [33], [41] if we specify a state
like φ, ψ or a number like 1,2 in a box in the YD then we call it the Young
Tableau (TY). So for states given above we give the corresponding Young
Tableau as
Young Diagram 143

1
φS ≡ 1 2 φA ≡ 2 (5.16)
So in this Young Tableau particle coordinates 1 and 2 specify the particle
in states φS and φA .
We saw here that the two S2 Young Diagrams were one-dimensional. By
building simple states φS and φA (due to the fact that this was the case with
just two particles) we could determine the dimensionality of its irreducible
representations. For arbitrary Sn to obtain the dimensions of irreducible rep-
resentation of Sn we follow the following rules.
When boxes in Young Diagrams are numbered with an integer, as we saw
it, is called the Young Tableau. We place numbers in Young Diagrams of Sn
as follows

(1) Numbers increase left-to-right along a row.

(2) Numbers increase on going down a column.

This is then denominated a standard tableau. The dimension of an irre-


ducible representation of Sn is equal to the number of standard tableau that
can be constructed from the corresponding partition.
A consistency check is provided by the formula:
X
li2 = n! (5.17)
i

where n! is the order of Sn and li are dimensions of the irreducible represen-


tation for a particular Young Diagram.

S2 : [2] ≡ 1 2 ; 1 − dimensional representation

1
[11] ≡ 2 ; 1 − dimensional representation

(check : 12 + 12 = 2 is 2! )

S3 : [3] ≡ 1 2 3 ; 1 − dimensional representation

1 2 1 3
[21] ≡ 3 , 2 ; 2 − dimensional representation
144 Group Theory in Particle, Nuclear, and Hadron Physics

1
2
[111] ≡ 3 ; 1 − dimensional representation

(Check 12 + 22 + 12 = 6 = 3!)

————————————————–
Problem 5.2: Obtain the dimensions of all the irreducible representations
of S4 .
————————————————–
Problem 5.3: Give the Young Diagram of partition of n = 5 for the
permutation group S5 and give their irreducible representations.
————————————————–

As is evident from above to find dimension of a particular irreducible


representation of Sn for large number of particles would be a tedious job. A
useful formula for finding the dimensions of an irreducible representation of
Sn is as follows:
n!
d[f ] = (5.18)
Πni hi
where [f ] is a given Young Diagram of Sn . In the denominator we have product
of hook lengths for each box in a Young Diagram. Hook lengths are defined
as follows. Hook length hi is a number (integer) associated with a particular
box ‘i’ in a YD. Take any box ‘i’, put a ‘dot’ at the centre of the box. Starting
from this central dot of the ith box draw an arrow cutting across all the boxes
on the right of the ith box. Next draw another arrow cutting across all the
boxes below it. The total number of boxes that these two arrows (including
the box with the dot in it) pass through is called the hook length of the ith
box in the YD. As an example take an arbitrary YD of S5 and take the second
box in the first row.

• →

So its hook length is 3. Next take the first box in the second row,

• →

Thus, its hook length is 2. Below we put hook length numbers for each
box of the YD as

4 3 1
2 1
Young Diagram 145

So using Equation 5.18 the dimensionality of the irreducible representation


of this YD of S5 is

5.4.3.2.1
d= 1.3.4.1.2 =5

As another example take the following Young Diagram of S4 . We label its


hook lengths inside the boxes.

4 1
2
1

4.3.2.1
d= 1.4.2.1 =3

————————————————–
Problem 5.4: Give all the Young Diagrams of S6 and then obtain their
dimensions using the hook-length formula.
————————————————–

Note that Sn the partition/Young Diagram [n] and [1n ](≡ [11, · · · 1] n −
number of 1’s) have one-dimensional irreducible representation each, while
all others have dimensions greater than one for their irreducible representa-
tions. Also Young Diagram/Tableau that are related by interchange of rows
and column are called conjugate representations. So the following Young
Diagram of S4 are conjugate representations.

and

Above we saw that the second YD has dimension three. The first one has
dimensions

4.3.2.1
d= 1.2.4.1 =3

So the conjugate diagrams have the same dimensions or the same number
of irreducible representations associated with them. Diagrams where the num-
bers of rows and columns are the same are termed self-conjugate diagrams.
So in S4 an example of a self-conjugate diagram is
146 Group Theory in Particle, Nuclear, and Hadron Physics

And its dimension is

4.3.2.1
d= 2.3.2 = 2.

5.2 Young Operators


We saw above that the subtle difference between the Young Diagram or
Young Tableau is that in the latter case numbers 1,2, .. exist inside each box.
These label the coordinates signifying all the degrees of freedom of identical
particle numbers 1,2,.. etc. But very often scientist do not make this distinction
– whether it is a Young Diagram or a Young Tableau is evident from the
context of the discussion.
We saw that φ(1, 2), a two-particle state and with permutation operator
P12 as φ(2, 1) = P12 φ(1, 2) and then with φS = φ(1, 2) + φ(2, 1) and φA =
φ(1, 2) − φ(2, 1), we have P12 φS = +φS and P12 φA = −φA . Let us define two
operators – symmetrizer S12 and antisymmetrizer A12 as

S12 = 1 + P12
A12 = 1 − P12 (5.19)
Then,

S12 φ(1, 2) = (1 + P12 )φ(1, 2) = φ(1, 2) + φ(2, 1) = φs

A12 φ(1, 2) = (1 − P12 )φ(1, 2) = φ(1, 2) − φ(2, 1) = φA (5.20)


And so the symmetrizer and antisymmetrizer acting on φ(1, 2), a state
with no specific/particular two particle symmetry, builds up a symmetric and
2
an antisymmetric state, respectively. Note P12 φ(1, 2) = P12 φ(2, 1) = φ(1, 2)
2
and so P12 = 1
Anticipating our use of exchange of arbitrary particle label numbers i ↔ j,
we define general symmetrizers and antisymmetrizers as

Sij = 1 + Pij
Aij = 1 − Pij (5.21)
Note,
Young Diagram 147

Sij Aij = (1 + Pij )(1 − Pij ) = 1 + Pij − Pij − Pij2 = 1 − 1 = 0

Let us now represent an n-particle state as a product of one particle states.

| α1 β2 γ3 · · · δn >=| α1 >| β2 >| γ3 > · · · | δn >


= α1 β2 γ3 · · · δn (5.22)
So the particle 1 in state α, particle 2 in state β ...etc.
To obtain the state with the required symmetry property for a Young
Tableau, we introduce a Young operator for each Young Tableau.
X X
Y =( Aν )( Sλ ) (5.23)
column rows

Note the first operator acting is the one on the right and the left one in
the second operation. Here

Sλ is the symmetrizer corresponding to λ boxes in a row.

Aν is the antisymmetrizer corresponding to ν boxes in a column.

The sums are taken over all rows and columns, respectively. Here sym-
metrizer is
X  1, 2, · · · λ 
Sλ = (5.24)
p1, p2 , · · · pλ
P
P
Here P is the sum over all the permutations in a row.
Antisymmetrizer is
X  1, 2, · · · ν 
Aν = P (5.25)
p1, p2 , · · · pν
P

P = + or − 1 corresponding to even or odd permutations.


The state with the symmetry of [ f ] standard Young Tableau is obtained
by

| φ[f ] >= Y[f ] | α1 β2 γ3 .......δn > (5.26)


Particularly important are one row only,

| φs >= Sn | α1 β2 .....δn> (5.27)


This corresponds to all the identical particles obeying Bose-Einstein statis-
tics (for spins S=0,1,2.. integral states).
And one column only,

| φA >= An | α1 β2 ........δn > (5.28)


148 Group Theory in Particle, Nuclear, and Hadron Physics

This corresponds to all the identical particles obeying Fermi-Dirac statis-


tics for spin − 12 , 32 , 52 ..., half integral spins.
Let us simplify and connect to our two particle state φ(1, 2)

φ(1, 2) =| α1 >| β2 >≡ α1 β2 (5.29)


Now  
1 2 , Y = AS = S = P 1, 2
P (note here A=1 operator)
p1, p2
 
P 1, 2
φS = Y α1 β2 = Sα1 β2 = P α1 β2
p1, p2
   
1, 2 1, 2
= α1 β2 + α1 β2
1, 2 2, 1
= α1 β2 + α2 β1 = α1 β2 + β1 α2 = αβ + βα

(using the convention that position labels 1,2 fixed and states exchanged.)
And,
1  
2 , Y = AS = A = P P 1, 2 (note here S=1 operator)
P p1, p2
 
P 1, 2
φA = Y α1 β2 = Aα1 β2 = P P α1 β2
p1, p2
   
1, 2 1, 2
= α1 β2 − α1 β2
1, 2 2, 1
= α1 β2 − α2 β1 = α1 β2 − β1 α2 = αβ − βα

As an example let us build all the states of S3 , the permutation group of


three particles. Its order is 3! = 6 and as we have seen, its elements are
     
1 2 3 1 2 3 1 2 3
ρ0 = , ρ1 = , ρ2 =
1 2 3 2 3 1 3 1 2
     
1 2 3 1 2 3 1 2 3
µ1 = , µ2 = , µ3 =
1 3 2 3 2 1 2 1 3
{ρ0 , ρ1 , ρ2 } are even permutations and {µ1 , µ2 , µ3 } are odd permuta-
tions. Let us take the totally symmetric state,

1 2 3 → ψs = S123 {α1 β2 γ3 } (Y = AS, A = 1 here)


 
P 1 2 3
= P (α1 β2 γ3 )
p1 p2 p3
     
1 2 3 1 2 3 1 2 3
= (α1 β2 γ3 ) + (α1 β2 γ3 ) + (α1 β2 γ3 ) +
1 2 3 2 3 1 3 1 2
Young Diagram 149
     
1 2 3 1 2 3 1 2 3
(α1 β2 γ3 ) + (α1 β2 γ3 ) + (α1 β2 γ3 )
1 3 2 3 2 1 2 1 3
= α1 β2 γ3 + α3 β1 γ2 + α2 β3 γ1 + α1 β3 γ2 + α3 β2 γ1 + α2 β1 γ3

ψs = αβγ + βγα + γαβ + αγβ + γβα + βαγ (5.30)


The totally antisymmetric function

1
2  
3 → ψA = Y α1 β2 γ3 = Aα1 β2 γ3 = Σ ∈P 1 2 3
(α1 β2 γ3 )
P p1 p2p3
     
1 2 3 1 2 3 1 2 3
= α1 β2 γ3 + α1 β2 γ3 + α1 β2 γ3 −
 1 2 3  2 3 1   3 1 2
1 2 3 1 2 3 1 2 3
α1 β2 γ3 − α1 β2 γ3 − α1 β2 γ3
1 3 2 3 2 1 2 1 3
ψA = αβγ + βγα + γαβ − αγβ − γβα − βαγ (5.31)
For mixed representations:

1 2
3 → ψ1 = Y α1 β2 γ3 = A13 S12 (α1 β2 γ3 )

A13 S12 = [1 − (13)][1 + (12)] = 1 + (12) − (13) − (13)(12)

(13)(12) ≡ µ2 µ3 = ρ1 = (123) so A13 S12 = 1 + (12) − (13) − (123)

ψ1 = [1 + (12) − (13) − (123)]α1 β2 γ3

= α1 β2 γ3 + α2 β1 γ3 − α3 β2 γ1 − α3 β1 γ2

= αβγ + βαγ − γβα − βγα

1 2
3 → ψ2 = Y α1 β2 γ3 = A23 S12 (α1 β2 γ3 )

A23 S12 = [1 − (23)][(1 + 12)] = 1 + (12) − (23) − (23)(12)

= 1 + (12) − (23) − (132)

ψ2 = [1 + (12) − (23) − (132)](α1 β2 γ3 )

= α1 β2 γ3 + α2 β1 γ3 − α1 β3 γ2 − α2 β3 γ1
150 Group Theory in Particle, Nuclear, and Hadron Physics

= αβγ + βαγ − αγβ − γαβ

1 3
2 → ψ3 = A12 S13 α1 β2 γ3

A12 S13 = [1 − (12)][(1 + 13)] = 1 + (13) − (12) − (12)(13)

ψ3 = [1 + (13) − (12) − (132)](α1 β2 γ3 )

= α1 β2 γ3 + α3 β2 γ1 − α2 β1 γ3 − α2 β3 γ1

= αβγ + γβα − βαγ − γαβ

1 3
2 → ψ4 = A23 S13 α1 β2 γ3

A23 S13 = [1 − (23)][(1 + 13)] = 1 + (13) − (23) − (23)(13)

ψ4 = [1 + (13) − (23) − (123)]α1 β2 γ3

= α1 β2 γ3 + α3 β2 γ1 − α1 β3 γ2 − α3 β1 γ2

= αβγ + γβα − αγβ − βγα

However the functions ψ1 , ψ2 , ψ3 , ψ4 are not orthogonal to one another.


Take, ψ20 = ψ2 − 12 ψ4 linear combination. Then 12 ψ1 and √13 ψ20 are orthogonal
two-dimensional irreducible representation of S3 . Using Schmitt Orthogonal-
ization process.

ψ30 = ψ3 + 14 ψ1 − 12 ψ20

And,

ψ40 = ψ4 − 14 (ψ1 + 2ψ20 + 2ψ30 )

√1 ψ 0
and 23 ψ40 form an orthogonal basis as giving additional two-
3 3
dimensional irreducible representations.

————————————————–
Problem 5.5: Write the Young operator for the following Young Dia-
grams for S4 and write the wave function represented by each tableau.
Young Diagram 151

1 2 3 1 2 4 1 3 4 1 2 1 3
(a) 4 (b) 3 (c) 2 (d) 3 4 (e) 2 4

————————————————–

5.3 Young Diagram and Unitary Symmetry


In quantum mechanics we represent the spin up and spin down doublet of
SU(2) group by the states,
   
1 1 −1 0
↑= χ 21 = and ↓= χ 1 2 = (5.32)
2 0 2 1
1
So there are these two states of a single particle of a spin 2 state χS .
We use a single box in Young Diagram to represent this one particle
1
state. So this stands for the spin 2 state χs . Next we map
1 1

χ 21 ≡ 1 and χ 1 2 = 2 (5.33)
2 2

So the Young Tableau with numbers specifies the spin coordinates of the
two spin states.
Next consider two such particle states. In order to be a multiplet (i.e.,
an irreducible representation) of the symmetric group S2 , the state must be
either symmetric corresponding to Young Diagram or antisymmetric

corresponding to Young Diagram .


If both particles are in spin up states (↑↑), then this is represented as
1 1 . and if both are in spin down state (↓↓), then that is represented as
2 2 . The other possible symmetry in the standard arrangement is 1 2 .
The nonstandard notation 2 1 is taken the same as the above and hence
not independent. So the symmetric two particle state is a triplet. Thus
the multiplicity is given by all the standard arrangements of Young Diagram
with integers restricted to 1 and 2. The only antisymmetric two particle
1
state is 2 which is a singlet. Thus,

2⊗2=3⊕1 (5.34)
152 Group Theory in Particle, Nuclear, and Hadron Physics

1 2
Note 1 or 2 are zero. So for the unitary group we modify the standard
arrangement rules for SU (n) as follows

(1) The numbers in Young Tableau do not decrease in going from left to
right.

(2) The numbers increase down a column.

Next for S3 the partitioning of n = 3 gives three Young Diagrams

[3] ≡

[21] ≡

[111] ≡

Now take three spin 12 states. With only two spin 21 states it is not possible
to build a totally antisymmetric state of three particles, and thus the contri-
bution of Young Diagram [111] above is zero. This property on what S3 allows
but what SU (2) does not allow imposes the extra constraint when applying
S3 to SU (2). Now only the first two Young Diagrams are applicable for three
spin 12 particles. As there are three particles but each is in only two states – ↑
or ↓ call it ↑≡ 1, ↓≡ 2. The standard Young Tableau for three spin 12 particle is

1 1 1 1 1 2 1 2 2 2 2 2

which is a quadruplet. Tableau of mixed symmetry,

1 1 1 2
2 3

This is a doublet. This is a doublet just like is.


Why is it so? This is so as there is only one way of making an antisymmetric
state of two particles which itself can exist in only two states (like spin-up and
spin-down). So we find that for SU (2) only columns of maximum two boxes
can exist. Each column of two boxes for SU (2) saturates. This word means
that for this state the spin is zero. So one obtains spin zero for each pair of
electrons (spin 21 ) in each column. Any extra electron exists on top of this new
“vacuum” background. So as far as the dimension of the representation goes,
Young Diagram 153

we can ignore the saturated boxes and just count the extra box sitting on top
of this inert background.
Now if we have the group SU (n) then in the fundamental representation
there are n-number of states and so a particle which has SU (n) as a degree of
freedom can be labelled by numbers 1, 2, ..n. We already saw that electron as
per SU (2) group can be labelled by two spin states 1 and 2. A quark which
forms the fundamental representation of the group SU (3) can therefore be
labelled by numbers 1, 2, and 3.
There is an important connection relating to Sα , the permutation group
of α− number of particle. This Sα has irreducible representation of Sα given
by all the allowed Young Diagrams as per the partitioning of the number
α. Let these identical particles correspond to the n-dimensional irreducible
representation (fundamental representation) of SU (n). Then it turns out that
all these YD of Sα also provide irreducible representations of the multiparticle
states obeying SUn group structure.
We saw this for the above spin 21 cases for two and three electrons. Now
for spin 21 one can build a maximum one antisymmetric state, i.e., one column
with two boxes, and a column of 2 + 1 = 3 boxes or more is not allowed. We
can have however three or more boxes in a row.
Similarly for SU (n) group for Sα , i.e., α number of particles, only maxi-
mum boxes α = n are allowed in a column in a Young Diagram. More boxes
then n in a column give zero. We can only build one antisymmetry state for
SU (n) with n boxes in a column.
The standard Young Tableau for three particles as per S3 is given as

1
1 2 1 3 2
1 ⊗ 2 ⊗ 3 = 1 2 3 ⊕ 3 ⊕ 2 ⊕ 3

The same Young Diagram of S3 translates one-into-one to be a valid Young


Diagram for three particles represented by the unitary group SU (n) (where
n ≥ 3) and so the Young Diagram for SU (n) for three particles is

⊗ ⊗ = ⊕ ⊕ ⊕

The same should hold for SU (2) with the constraints that the last one
could give zero. So we know that if we take three spin 21 electrons then the
final states of good spins are one spin 32 and two spin 21 states. Thus we have
from above
154 Group Theory in Particle, Nuclear, and Hadron Physics

⊗ ⊗ = ⊕ ⊕

2⊗2⊗2=4⊕2⊕2

where 4 is the 4-dimensionality of spin 32 states (+ 23 , 12 , − 12 , − 32 ) and two spins


10
2 s. Note that the total dimensionality on the left 2 × 2 × 2=8 is equal to the
ones on the right 4+2+2=8.
We now formalize the working definition as to the dimensions of the irre-
ducible representation of SU (n) as follows. Given a Young Tableau of SU (n),
the dimensionality of irreducible representation is given by standard arrange-
ment as

1. Put an integer from 1 to n in each box of YT such that the number


increases from top to bottom in a column.

2. These numbers do not decrease from left to right in a row.

So the numbers have to be all different in a column but may be the same
in a row.

Now
denotes a single particle state of SU (n). Here numbers 1 to n can be put

in the box as a standard arrangement. Next two particle states and .

We saw what these were for SU (2). Now for SU (3) the standard arrange-
ments yield the following dimensionality:

1 1 1 2 1 3 2 2 2 3 3 3 ; 6-dimensions

1 1 2
2 3 3 ; 3-dimensions

Next three particle states,

⊗ ⊗ = ⊕ ⊕ ⊕

1 1 1 1 1 2 1 1 3 1 2 2 1 2 3 1 3 3 2 2 2
Young Diagram 155

2 2 3 2 3 3 3 3 3 ; 10-dimensions

1 1 1 1 1 2 1 2 1 3 1 3 2 2 2 3
2 3 2 3 2 3 3 3 ; 8-dimensions

1
2
3 ; singlet

3 ⊗ 3 ⊗ 3 = 10 ⊕ 8 ⊕ 8 ⊕ 1

Now 3 × 3 × 3=27 on the left side is equal to 10+8+8+1=27 on the right.

Next in SU (5) say we have a Young Tableau,

As the column with 5 boxes saturates for SU (5), its dimensionality is the
same as that of the unsaturated tableau,

Note that the first one is a 16-particle state, and the second one is a 6-
particle state but the dimensions of the two are exactly the same. So if we are
only interested in the dimensions of the above 16-particle state, we can ignore
the saturated boxes in the first two columns and obtain the dimension for the
simpler and smaller tableau – the second one.
By the above method we obtain the dimensions of an irreducible repre-
sentation of a YD of SU (n) [42], [32]. But it gets tedious for a large number
of particles and a large value of n in SU (n). There exists a useful graphical
method to determine the above dimensions in a convenient manner.
For SU (n) define a “Distance of the first box (Di )”. This is the number of
steps going from the box in the left-hand corner of the Young Diagram (i.e.,
the first box) to the ith box with each step to the right taken as +1 and each
downward steps along a column as -1. So Di for the following Young Diagram
is
156 Group Theory in Particle, Nuclear, and Hadron Physics

0 1 2
-1 0
-2

Next is “hook length” hi as defined earlier for the YD of Sn . This hook


length hi for the YD are given as the numbers in the boxes,

5 3 1
3 1
1

Thus the dimension of the SU (n) irreducible representation associated


with a Young Tableau is

Πi (n + Di )
d= (5.35)
Πi hi
So for SU (n) the YD is,

0 1 3 1
the distances are -1 and the hook lengths are 1

So for SU (n) for this diagram dimension is


n(n+1)(n−1) n(n+1)(n−1)
d= 1.3.1 = 3

And for SU (3) d = 3×4×2


3 = 8.
Now for the seven box YD of SU (n),

The hook lengths are

6 5 3 2 1
2 1

And the distances are

0 1 2 3 4
-1 0

3.4.5.6.7.2.3
For the SU (3) group the dimension of this YD is, d = 1.2.3.5.6.2.1 = 42
Young Diagram 157
2.3.4.5.6.1.2
And for the SU(2) group the dimension of this YD is, d = 1.2.3.5.6.2.1 =4

Thus the dimensionality corresponding to a Young Diagram is different


for different groups. However the denominator (hook length) is independent
of SU (n) and depends just upon the number of boxes in a particular Young
Diagram.

————————————————–
Problem 5.6: For SU (3) consider the Young Diagrams for three particles
and obtain their dimensions and confirm that 3 ⊗ 3 ⊗ 3 = 10 + 8 + 8 + 1.
————————————————–

We have been specifying the Young Diagrams by giving their dimensions.


Another technique is commonly used to specify a Young Diagram for SU (n)
[32]. The irreducible representation for SU (n) is specified by giving (n − 1)
integers (p1 , p2 , .......pn−1 ) as

p1=λ1 −λ2

p2=λ2 −λ3

··

pn−1=λn−1 −λn (5.36)


So for SU (3) the YD is specified by 3-1=2 integers ( p1 , p2 ) as follows:

(0, 1)

(2.0)

(1, 1)

(2, 1)

(2, 2)
158 Group Theory in Particle, Nuclear, and Hadron Physics

Given a Young Tableau specified by (p1 , p2 ....pn−1 ) for SU(n). Its conju-
gate tableau is specified by (pn−1 , pn−2 ......p1 )’ So for SU (3) given YD,

λ1 = 3, λ2 = 2 ; p1 = 1, p2 = 2 ; i.e.(1, 2)

So the conjugate tableau is (2,1) as,

λ1 = 3, λ2 = 1 ; p1 = 2, p2 = 1 ; i.e.(2, 1)

Now the dimension of (1,2) YD is

3.4.5.2.3
d= 1.3.4.1.2 = 15

and that of the conjugate YD (2,1) is

3.4.5.2.
d= 1.2.4.1 = 15

Hence the two conjugate YD have the same dimensions. Also note that
the conjugate diagrams have different numbers of boxes. In contrast what
we defined as a conjugate diagram in Sn the permutation group, where the
conjugate diagram had the same number of boxes. So one should specify which
conjugate diagram one is talking about.
The two conjugate diagrams of SU (n) are useful in the sense that when
appropriately placed along the original YD, like a jigsaw puzzle, create a rect-
angle of n rows for SU (n). So the above conjugate diagram put together with
the first YD creates the following 3 × 3 rectangle.

X
X X X

Note that the part with the X’s is the conjugate diagram inverted to fit
the original YD. The whole now is a saturated YD.
So the dimensionality of the conjugate YD though the same corresponds to
two different inequivalent representations. For SU (2) these are still equivalent
representations as these are specified by one number p1 . Therefore the 2-
representations of SU (2) is termed the pseudo-real representation.
A Young Diagram is self-conjugate if it coincides with its conjugate. So for
SU (3),
Young Diagram 159

is self-conjugate.

As we saw, a diagram and its conjugate give saturated columns n-boxes.


Now for SU (n) we have from one-box to (n − 1)-boxes in a single column. All
these correspond to what we call as the fundamental representations of
SU (n). The fundamental representations are the irreducible representations
which given (p1 , p2 , .....pn−1 ) is restricted to pi = 1; pj = 0; j 6= 1. So SU (n)
has (n-1) fundamental representation.

1. SU (2), , 2-dimensional – just one fundamental representation.

2. SU (3), , 3-dimensional, , 3̄-dimensional – bar on 3 to indicate


inequivalent representation from 3, so 2-fundamental representations.
These two fundamental representations we take as corresponding to the
quark and the antiquark, respectively, in SU (3)F , the flavour group in the
quark model, or as standing for colour and anticolour in the SU (3)C , the
colour group of QCD.

3. SU(4), , 4-dimensional, , 6-dimensional, , 4̄-dimensional –


bar on 4 to indicate difference from 4, so 3-fundamental representations.

Now given SU (n), with the fundamental representation


, its conjugate representation has (n − 1) boxes in the column.

For SU (3):

3 ⊗ 3̄ = 1 ⊕ 8

For SU (n) this generalizes to,

n ⊗ n̄ = 1 ⊕ (n2 − 1).
160 Group Theory in Particle, Nuclear, and Hadron Physics

5.4 Product of Irreducible Representations


If we are taking a product of two Young Diagrams then in the second one
on the right, put integers 1,2,3,..n-1 (for SU (n)) as numbering the different
rows. So,

1 1 1 1
2 2
3

Then enlarge the first YD (on the left) by enlarging with the various boxes
from the second diagram one by one, with the following conditions being ob-
served in each step.

(a) Each diagram must be a proper Young Diagram - meaning each row
below not larger than the one above it and columns having only up to n-boxes.

(b) Numbers in a row should not decrease from left to right. In a column
all numbers should be different from each other and increase top to bottom.

(c) Finally count all symbols right to left down each row, the number ‘i’
must at no time have occurred more often than the number (i − 1). So (11213)
is all right but (12211) is forbidden.

So for SU (3),

⊗ 1 = 1 ⊕ 1 ; 3 ⊗ 3 = 3̄ ⊕ 6

⊗ ⊗ =( ⊕ )⊗ 1

= ⊗ 1 ⊕ ⊗ 1

1
=( 1 ⊕ )⊕( 1 ⊕ 1 )
Young Diagram 161

=( ⊕ )⊕( ⊕ )

3 ⊗ 3 ⊗ 3 = 1 ⊕ 8 ⊕ 8 ⊕ 10

These check out. This is called the reduction of the product of irreducible
representations.
Take product

1 1 1
⊗ 2 → ⊕ 1 ⊕ 1

1 1
1 1 1 1 1
→( ⊕ 1 ⊕ 1 )⊕( 1 ⊕ 1 )⊕( 1 ⊕

1
1 )

1 1 1 1
1 1 1 1 2 1
→ 2 ⊕ 2 ⊕ 1 2 ⊕ 2 ⊕ 1 ⊕ 1 2

⊗ =•⊕ ⊕ ⊕ ⊕ ⊕
(5.37)

¯ ⊕ 27
8 ⊗ 8 = 1 ⊕ 8 ⊕ 8 ⊕ 10 ⊕ 10 (5.38)

≡ • − singlet state

¯
; d = 10

; d = 27
162 Group Theory in Particle, Nuclear, and Hadron Physics

Note that 8 ⊗ 8 gives a singlet plus other states. This singlet plays an
important role in the quark model.
Note that as representation (300) of dimension 10

and ¯
as representation (003) of dimension 10

are conjugate diagrams. The bar on one indicates that it is an independent


representation to the state with dimension 10. Also it may be treated as anti-
state of the other state.
These are conjugate to each other as putting these two YD together as in
a jigsaw puzzle, we get a saturated box diagram.

X X X
X X X

————————————————–
Problem 5.7: For SU (n) obtain n ⊗ n̄ states.
————————————————–

Problem 5.8: Find (a) ⊗ (b) ⊗


————————————————–

5.5 Multiplets of the SU (n − 1) Subgroup of SU (n)


The fundamental representation of SU (3) is a triplet and this contains a
doublet plus a singlet of SU (2). We wish to identify the SU (2) multiplets in
the octet state of SU (3). The eight standard arrangements for the octet state
in SU (3) are

1 1 1 1 1 2 1 2 1 3 1 3 2 2 2 3
2 3 2 3 2 3 3 3

Now 3 refers to the third entity on top of the (1,2) of SU (2). So in the
above,
Young Diagram 163

1 1 1 2
2 , 2 → 1 , 2 → , this corresponds to the SU (2) doublet.

Now in above YT, 3 occurs only once in,

1 1 1 2 2 2 1 3
2 , 3 3 , 2

By removing 3 this reduces to the SU (2) diagrams,


1
(1 1 , 1 2 2 2 ), 2

The three in bracket gives the SU (2) triplet and the 4th one is the SU (2)
singlet. The remaining two diagrams have 3 occurring twice.

1 3 2 3
3 , 2 → 1 , 2 → , this corresponds to the SU (2) doublet.

So the octet of SU (3) has 2 SU (2) doublets, a singlet and a triplet. This
decomposition is expressed as

8⊃1⊕2⊕2⊕3 (5.39)

In compact manner the same is seen as follows by just filling in only the 3’s.

3 3
→ ⊕ ⊕ 3 ⊕ 3
SU (2)
8 −−−−→ 2 ⊕ 1 ⊕ 3 ⊕ 2

By generalizing, we can find SU (n − 1) content of multiplets contained in


a given Young Tableau of SU (n). The trick is to consider all possible ways of
putting n0 s in boxes consistent with the standard arrangement of the tableau
of SU (n). Number n may occur zero number of times, once, twice, etc. con-
sistent with the standard arrangement of the tableau. Next remove the boxes
with n’s in them and what remains is the SU (n − 1) multiplet content. So for
example,

n n n n
→ ⊕ ⊕ n ⊕ ⊕ n ⊕
164 Group Theory in Particle, Nuclear, and Hadron Physics

n n
n

→ ⊕ ⊕ ⊕ ⊕ ⊕
n(n+1)(n+2)(n−1)
The SU (n) dimension of YD on left is, d = 2.4 =
2
n(n −1)(n+2)
8

(n−1)(n)(n+1)(n−2)
The SU (n − 1) dimension of the first YD is d = 8 =
2
n(n −1)(n−2)
8

(n−1)n(n−2) n(n−1)(n−2)
second YD, d = 3 = 3

(n−1)n(n+1) (n2 −1)n


third YD is, d = 2.3 = 6

(n−1)(n−2) (n−1)(n−2)
fourth YD is, d = 2 = 2

(n−1)n n(n−1)
fifth YD is d = 2 = 2

sixth YD is, d = n − 1
n(n2 −1)(n+2) 2
−1)(n−2) 2

8 ⊃ n(n 8 + n(n−1)(n−2)
3 + n(n6−1) + (n−1)(n−2)
2 + n(n−1)
2 +
(n − 1)

Check that the total of all the multiplicities of the RHS is that of the LHS.
In the same manner we can determine SU (n − 2) multiplets of SU (n − 1) and
so on.

————————————————–
Problem 5.9: Given the following Young Tableau for the group SU (4),
determine its reduction with respect to SU (3).

————————————————–
Problem 5.10: Find the SU (2) decomposition of the decuptet of SU (3).
————————————————–
Young Diagram 165

5.6 The Reduction SU (m + n) → SU (m) ⊗ SU (n)


Consider the following under the group reduction SU (m + n) → SU (m) ⊗
SU (n)

m + n → (m, 1) ⊕ (1, n) (5.40)


So under SU (m + n) → SU (m) ⊗ SU (n) the above breakup in terms of
Young Diagram is

→( , •) ⊕ ( • , ) (5.41)
Remember the representation with no boxes is the identity representation,
which means that it is a singlet of dimension 1 and given by •. Also remember
that a column with N boxes for SU (n) is a singlet too.
We build up product representations of SU (m + n) step-by-step while
performing the reduction as above. So,

⊗ → [( , •) ⊕ ( • , )] [( , •) ⊕ ( • , )] (5.42)

(m + n) ⊗ (m + n) → [(m, 1) ⊕ (1, n)] ⊗ [(m, 1) ⊕ (1, n)]

( , •) ⊗ ( , •) ⊕ ( , •) ⊗ ( • , )⊕( • , )×( , •) ⊕ ( • , )⊗( • , )


(5.43)

(m, 1)(m, 1) ⊕ (m, 1)(1, n) ⊕ (1, n)(m, 1) ⊕ (1, n)(1, n)

[( , •) ⊕ ( , •)] ⊕ ( , ) ⊕( ) ⊕ [( • , ) ⊕ (• , )]
(5.44)

{ 12 m(m + 1), 1} + { 12 m(m − 1), 1} + (m, n) + (m, n) + {1, 21 n(n + 1)} +


{1, 12 n(n − 1)}
166 Group Theory in Particle, Nuclear, and Hadron Physics

So under the reduction SU (m + n) → SU (m) ⊗ SU (n)

→( , •) ⊕ ( , ( ) ⊕ (• , ) (5.45)
1
2 (m + n)(m + n + 1) → { 12 m(m + 1), 1} + {m, n} + {1, 12 n(n + 1)}

and

→( , •) ⊕ ( , ( ) ⊕ (• , ) (5.46)
1
2 (m + n)(m + n − 1) → { 12 m(m − 1), 1} + {m, n} + {1, 12 n(n − 1)}

check 12 (m + n)(m + n + 1) ⊃ 12 m(m + 1) + mn + 21 n(n − 1)

= { 12 m(m + 1) + 12 mn} + { 12 mn + 12 n(n + 1)}

= 12 m(m + n + 1) + 12 n(m + n + 1) = 21 (m + n)(m + n + 1)

Similarly 12 (m + n)(m + n − 1) ⊃ 12 m(m − 1) + mn + 12 n(n − 1)

Note that taking n = 1 then this should give l = m + 1 reduction to (l − 1)


for SU (l) ⊃ SU (l − 1) as above.

For the three particle states, we have,

→( , •) ⊕ ( , )⊕( , ) ⊕ (• , )
(5.47)
and,

→( , •) ⊕ ( , )⊕( , ) ⊕ (• , ) (5.48)

————————————————–
Unsolved Problem 5.1:
Confirm above Equations 5.47 and 5.48.
————————————————–

As another application of this consider the extension from the three-flavour


(u, d, s) quark SU (3)F model to four-flavour (u,d,s,c) quarks in SU (4)F . Let
us study the reduction SU (4)F → SU (3)F ⊗ U (1)c where U (1)c stands for
Young Diagram 167

charm c-quark. (u, d, s, c) forms the fundamental representation of SU (4)F and


(u,d,s) forms the fundamental representation of SU (3)F . So under SU (4)F →
SU (3)F ⊗ U (1)c , 4 → 3o + 1 where in subscript we give c-quark content of
the representation. So (u,d,s) or 3 has no c-quark and so is 3o , and 1 is one
c-quark and so representation 1, has one c-quark.
As seen above,

→( , •) ⊕ ( • , )
Now two quark state of SU (4)F ,

⊗ = ⊕
4 ⊗ 4 = 10 + 6

From above their reductions are into SU (4)F → SU (3)F ⊗ U (1)c .

→( , •) ⊕ ( , ) ⊕ (• , )
10 → 60 ⊕ 31 ⊕ 12

and

→( , •) ⊕ ( , ) ⊕ (• , )
6 → 3̄0 ⊕ 31
The last one being zero.
The three quark baryon in SU (4)F correspond to,

⊗ ⊗ = ⊕ ⊕ ⊕
4 ⊗ 4 ⊗ 4 = 20S ⊕ 20M S ⊕ 20M A ⊕ 4̄

We use S, MS, MA to distinguish the three states.

The reduction of these states under SU (4)F → SU (3)F ⊗ U (1)c is,

→( , •) ⊕ ( , )⊕( , ) ⊕ (• , )

20S → 100 ⊕ 61 ⊕ 32 ⊕ 13
168 Group Theory in Particle, Nuclear, and Hadron Physics

→( , •) ⊕ ( , )⊕( , )⊕( , )

20M → 80 ⊕ 61 ⊕ 3̄1 ⊕ 32

→( , •) ⊕ ( , )
4̄ → 10 ⊕ 3̄1

————————————————–
Problem 5.11: In the SU (6)SF model (SU (3)F and SU (2)S of
quark model are combined) the fundamental representation consists of
(u↑ , u↓ , d↑ , d↓ , s↑ , s↓ ) (in a column actually). We can break it up as
(u↑ , u↓ , d↑ , d↓ ) of SU (4)SF and (s↑ , s↓ ) SU (2)SF . Obtain the reduction of

and under SU (6)SF → SU (4)SF ⊗ SU (2)SF .


————————————————–
Problem 5.12: Now SU (6)F consists of (u, d, s, c, b, t) six flavor of quarks
as its fundamental representation. Three of these are light (u,d,s) and the
other three (c,b,t) are heavy in mass. So consider the breakup SU (6)F →

SU (3)L + SU (3)H and get reduction of and for it.


————————————————–

5.7 Reduction of SU (mn) → SU (m) ⊗ SU (n)


Under this reduction

→( , )
(mn) → (m, n)

Next two particle states


Young Diagram 169

⊗ →( , )⊗( , )

( ⊗ )→( ⊗ , ⊗ )
(mn)(mn) → (mm, nn)

=( ⊕ , ⊕ )

=( ⊕ )⊕( , )⊕( , )⊕( , )

Hence we take

→( , )⊕( , )

1
2 (mn)(mn + 1) → [ 12 m(m + 1), 12 n(n + 1)] + [ 12 m(m − 1), 12 n(n − 1)]

Check dimensions

1
4 mn(m + 1)(n + 1) + 14 mn(m − 1)(n − 1)

1
4 mn[(m + 1)(n + 1) + (m − 1)(n − 1)] = 12 mn(mn + 1)

and,

→( , )⊕( , )

1
2 (mn)(mn − 1) → [ 12 m(m + 1), 12 n(n − 1)] + [ 12 m(m − 1), 12 n(n + 1)]

→ 41 mn(m + 1)(n − 1) + 14 mn(m − 1)(n + 1) = 12 mn(mn − 1)

Now state in SU (mn) → SU (m) ⊗ SU (n) needs,

⊗ →( , )⊗( , )⊕( , )⊗( , )

Keeping the relevant diagrams we get


170 Group Theory in Particle, Nuclear, and Hadron Physics

→( , )⊕( , )⊕( , ) (5.49)

Now let us obtain of SU(mn). We need

→( , )⊕( , )

⊗ →( , )⊗( , )⊕( , )×( , )

→( , )⊕( , )⊕( , ) (5.50)

Taking reduction SU (6)SF ⊃ SU (3)F ⊗ SU (2)S , the non-zero terms give

→( , )⊕( , ) (5.51)
56S → (10, 4) ⊕ (8, 2)

And

→( , )⊕( , ) (5.52)
20A → (8, 2) ⊕ (1, 4)

These results are useful in the SU (6)SF quark model.

————————————————–
Problem 5.13: Use SU (mn) → SU (m)⊗SU (n) reduction for SU (8)SF ⊃
SU (4)F ⊗ SU (2)S i.e. (u, d, s, c) as four states of SU (4)F and (↑, ↓) as two
Young Diagram 171

of SU (2)S . Discuss its application for the baryons in SU (8)SF quark model
assuming that these are made up of three quarks.
————————————————–
Problem 5.14: Assume that out of the three quarks in the symmetric
SU (6)F S quark model, two are strongly bound as an s-state diquark. What
will the total baryonic spectrum look like?
————————————————–

5.8 Coefficients of Fractional Parentage and Isoscalar


Factors
We studied the permutation group of n objects Sn in detail earlier. We
took outer products of the permutation groups as this is relevant for studies
of the special unitary group SU (N ). The Young Diagram technique was a
powerful graphical technique which allows us to study how Sn and SU (N )
are related when various kinds of products are taken.
We saw that using a symmetric group in a tensor space, it was possible
to define basis vectors which form irreducible representation subspaces, both
under Sn and SU (n). The reduction SU (mn) ⊃ SU (m) ⊗ SU (n) as we see in
terms of Young Diagrams is actually related to the Inner/Kroenecker product
in the symmetric group. This involves Clebsch-Gordan (CG) coupling coeffi-
cients. The CG coefficients of SU (mn) are related to the CG coefficient of the
subgroups SU (m) and SU (n). The factors which bring out this relationship
are called Isoscaler Factors.
Given two irreducible representation [µ] and [µ0 ] of Sn . The Kroe-
necker/Inner/Tensor product [µ ⊗ µ0 ] is also a representation of Sn . How-
ever it is in general a reducible representation. It can be completely reduced
into a direct sum of irreducible representation components [µ00 ]. The two are
connected through the Clebsch-Gordan (CG) series.
X
[µ] ⊗ [µ0 ] = m(µ, µ0 , µ00 )[µ00 ] (5.53)

m(µ, µ0 , µ00 ) is the (inner) multiplicity index of the irreducible representation


[µ00 ] in the product [µ ⊗ µ0 ].
We have already seen for S3 an example of CG-series

[21] ⊗ [21] = [3] + [21] + [111]

For the tensor product space we defined two different orthonormal basis,
[µ⊗µ0 ] [µ00 ]
the product basis φM M 0 and the sum basis φM 00 (Γ). These two orthonormal
basis are related by unitary transformations. The matrix elements of this
172 Group Theory in Particle, Nuclear, and Hadron Physics

transformation are called Clebsch-Gordan (CG) coefficient. In Sn we choose


phases to ensure real CG coefficient (note Γ is a particular representation ).

[µ00 ]
X  µ µ0 µ00

[µ⊗µ0 ]
φM 0 (Γ) = C 0 00 φM M 0 (5.54)
0
M M M Γ
MM

Note the product state S = 1 as ↑↑, √12 (↑↓ − ↓↑), ↓↓. Here take S =
1 ∼ µ00 and ↑↑ etc. states as [µ ⊗ µ0 ]. The coefficient for ± √12 above are
CG coefficients for SU (2). Note that the intrinsic connection between Sn and
SU (n) is what allows to use the SU (2) states and still talk of representations
of S2 . We will study this deep connection below.
Note that the CG coefficients of the group Sn are related to the CG coef-
ficient of the group Sn−1 , as [25]


µ0 µ00 µ0 µ00 ν0 ν 00
    
µ X µ ν
Cn = Kn Cn−1
M M0 M 00 Γ
ν ν0 ν 00 α
K K0 K 00 α
α
(5.55)
νn
When ν = M n
, K = M1 · · · Mn−1 .
These K’s are called Isoscalar Factors. We will study these below.
The group Sn has amongst its subgroups the direct products Sn1 ⊗ Sn2
with n1 + n2 = n . In Sn if an element a of Sn is also an element of Sn1 ⊗ Sn2
then we can write it as a = a1 a2 when a1 ∈ Sn1 and a2 ∈ Sn2 . All the ir-
reducible representations of Sn1 × Sn2 are obtained by taking the products
[α] × [β] = [α × β] when [α] and [β] correspond to Sn1 and Sn2 , respectively.
But the corresponding matrices providing this representation do not form a
representation of Sn and it is not defined for those elements of Sn which are
not present in the subgroup Sn1 × Sn2 . However it is possible to induce a rep-
resentation of Sn and is called the outer product of irreducible representation
of [α] of Sn1 and [β] of Sn2 and is denoted as {α × β}. The outer product in
general can be completely reduced into its irreducible component through its
Outer-Product (OP) series.
X
{α × β} = n(α, β, µ)[µ] (5.56)
µ

Here n(α, β, µ) is outer-multiplicity index of the irreducible representation


[µ] into the induced representation {α × β}.
Note the difference between the CG-series and the OP-series equation.
In the CG series the irreducible representation [µ],[µ0 ], [µ00 ] all refer to the
same group Sn . However, in the OP- series the irreducible representation
[α], [β]and[µ] belong to different series Sn1 , Sn2 and Sn , respectively.
The rules for obtaining these reductions are very similar to the ones we used
to obtain the irreducible representation of product earlier. The only difference
is that however in Sn we can have more number of boxes in a column than
Young Diagram 173

restricted by SU (N ) - N being maximum. So the outer product of {[21]×[21]}


for S3 × S3 ⊂ S6 as follows:

1 1 1
1 1 1 1 1 2
⊗ 2 = 2 ⊕ 2 ⊕ 1 2 ⊕ 1

1
1 1
1 1 1 1
⊕2 ⊕ 2 ⊕ 1 2 ⊕ 2
(Note the difference with respect to our earlier diagrams where all these
were used except the four-column ones which are zero for SU (3)).
Similarly we can define product of irreducible representations [α] and [β]
of SU (N ). These are given as,
X
[α × β] = n(α, β, µ)[µ] (5.57)
µ

where n(α, β, µ) is the outer multiplicity index defined in eqn. (56) for Sp . For
the corresponding dimensions we have,

X
N [α]N [β] = n(α, β, µ)N [µ] (5.58)
µ

Thus (as already seen and used above), the Young Diagram or partition of
the CG-series of SU (N ) is exactly the same as the OP-series for Sn . There is
one difference that when n(α, β, µ) 6= 0 the Sn irreducible representation [µ]
appears in the OP-series. However in SU (N ), N [µ] = 0 when Young Diagram
[µ] has more than N rows. We pointed this out above, in the product of
{[21] × [21]} vis-a-vis that of the group SU (3).
The connection between the OP-series for Sn and the CG-series for SU (N )
can be understood as follows. Given Young Diagram [α] and [β] with boxes
n1 and n2 , respectively, then the irreducible representation of general tensors
X [α] and X [β] have permutation symmetries determined by Sn1 and Sn2 . The
symmetry of the product {α × β} is than described by Sn1 × Sn2 . The irre-
ducible representation of term X [α×β] should have a permutation symmetry
determined by Sn . And this is connected by the OP-series for Sn ⊃ Sn1 × Sn2 .
Now let us determine how to choose the orthonormal basis vectors belong-
ing to an irreducible representation [µ] of SU (N ) . One fruitful method is to
choose the basis vectors in a manner that these are eigenvectors of a set of
mutually commuting operators. For SU (N ) these are Casimir operators and
their weights.
For SU (2) we choose the Jz operator whose eigenvalue for a j takes the
174 Group Theory in Particle, Nuclear, and Hadron Physics

value −j to +j. so the SU (2) irreducible representations are specified by the


labels m = {I, I3 }.
For SU (3) we need three eigenvalues. Using decomposition SU (3) ⊃
SU (2) × U (1) we can choose two of the three operators in SU (2). Thus,

m = {I, I3 , Y }

Where Y is the hypercharge, the eigenvalue of the second diagonal gener-


ator of SU (3).
Using the decomposition SU (N ) ⊃ SU (N − 1) × U (1) for group
SU (N ), N > 2.

m = {[µ0 ], m0 , m00 }

where [µ0 ] and m0 are SU (N − 1) labels and m00 is the eigenvalue of the re-
maining operator. Thus for SU (4)F one may write it as,

m = {[µ0 ], I, I3 , Y, C}

where [µ0 ] is the SU (3) irreducible representation, I, I3 , Y are the SU (3) ”mag-
netic quantum numbers”, and C is charm, an additive quantum number re-
lated to the third diagonal generator of SU (4).
The above subgroup decomposition SU (N ) ⊃ SU (N − 1) × U (1) is not
the only way to generate m. We may have SU (3) ⊃ SO(3) or SU (4) ⊃
SU (2) × SU (2) etc.
Now this connection between SU (N ) and Sn is comprehensively given
by [25]. It turns out that there exists a direct relation between the Coeffi-
cient of Fractional arentage (CFP) related to the unitary group decomposition
SU (mn) ⊃ SU (m) ⊗ SU (n) and the Isoscalar Factor (ISF) arising from the
permutation group chain Sf1 +f2 ⊃ Sf1 × Sf2 .
We take S(f1 ) and S(f2 ) as the permutation groups for the particles 1, 2,
....f1 and f1 + 1, f2 + 2, .... f1 + f2 , respectively, with f = f1 + f2 . We use
Chen’s notation to specify the irreducible representations of the nine groups
S x (f1 ).....S q (f ) and the irreducible representations of SU (m), SU (n) and
SU (mn) ( with SU (mn) ⊃ SU (m) × SU (n) ) [43]
σ0 µ0 ν0
 x
S y (f1 ) S q (f1 )
  
S (f1 ) SU (m) SU (n) SU (mn)
 σ 00 µ00 ν 00   S x (f2 ) S y (f2 ) S q (f2 )   SU (m) SU (n) SU (mn) 
σ µ ν S x (f ) S y (f ) S q (f ) SU (m) SU (n) SU (mn)
(5.59)
{σ}, {µ} and {ν} are partition labels for the irreducible representations
of SU (m), SU (n) and SU (mn), respectively. So, {µ0 } and {σ 0 } are the irre-
ducible representations of S x (f1 ) and S y (f1 ), respectively. Let

0
{ν 00 }
  
0 0 {ν0 } 0 0 {ν}

β {σ }X1 {µ }X2 , β 00 {σ 00 }X100 {µ00 }X200 , β{σ}X1 {µ}X2 (5.60)

Young Diagram 175

be the SU (mn) ⊃ SU (m) × SU (n) irreducible bases in the q-space of the


particles (1, 2, · · · f1 ), (f + 1, · · · f ) and (1, 2, · · · f ), respectively. X1 (X2 ) is the
component index for the irreducible representation {σ} ({µ}); and β 0 , β 00 and
β are inner multiplicity labels:
β 0 = 1, 2, · · · (σ 0 µ0 ν 0 ), β 00 = 1, 2, · · · (σ 00 µ00 ν 00 ), β = 1, 2, · · · (σµν).
(σ 0 µ0 ν 0 ), (σ 00 µ00 ν 00 ) and (σµν) integers are determined by the CG series of the
permutation group. The SU (mn) ⊃ SU (m) × SU (n) CFP are defined as the
expansion coefficients as follows:


{ν}τ X {ν}τ,β{σ}θ{µ}φ

βσX1 µX2 = Cν 0 β 0 σ0 µ0 ,ν 00 β 00 σ00 µ00
β 0 β 00 σ 0 σ00µ0 µ00 θφ
{σ}θ{µ}φ
{ν 0 } {ν 00 }
 
0 0 0
βσµ
00 00 00
β σ µ (5.61)
X X 1 2

0 00
where τ , θ and φ are the outer multiplicity labels τ = 1, 2, · · · {ν ν ν}, θ =
0 00 0 00 0 00 0 00 0 00
1, 2, · · · {σ σ σ}, φ = 1, 2, · · · {µ µ µ}. {ν ν ν}, {σ σ σ} and {µ µ µ} are
integers which are determined by the Littlewood rule. Finally the bases are
combined into the irreducible basis {σ}X1 and{µ}X2 in terms of the CG
coefficients of SU (m) and SU (n), respectively.
Now the irreducible representation bases provide us the irreducible repre-
sentation of the group decomposition S(f ) ⊃ S(f1 ) × S(f2 ) in the x, y and
q(x, y) spaces are given as

  
{σ} {µ} {ν}
0 0 00 00 ,
0 0 00 00 ,
0 0 00 (5.62)
θ{σ }m {σ }m
1 1 φ{µ }m2 {µ }m2 τ {ν }m {ν 00 }m
0 0
here m1 , m2 , etc. are the Yamanouchi symbols. The S(f ) ⊃ S(f1 ) × S(f2 )
ISF are given as the expansion coefficients as


{ν}β X 0
{ν}β,τ {ν }β {ν }β
0 00 00

τ ν 0 m0 ν 00 m00
= C{σ}θσ0 σ00 ,{µ}φµ0 µ00
β 0 β 00 σ 0 σ 00 µ0 µ00 θφ
  {ν 0 }β 0 {ν 00 }β 00
{σ} {µ}
θσ 0 σ 00

φµ0 µ00
(5.63)
m0 m00

Note that the coupling in terms of the CG-coefficients is indicated as su-


perscripts and subscripts on the square bracket above.
It is important to note that the SU (mn) ⊃ SU (m) × SU (n) CFP are
identical with the S(f ) ⊃ S(f1 ) × SU (f2 ) ISF, i.e.,
0 0 00 00
{ν}τ,β{σ}θ{µ}φ {ν}β,τ {ν }β {ν }β
C{ν 0 }β 0 σ0 µ0 ,{ν 00 }β 00 σ00 µ00 = C{σ}θσ0 σ00 ,{µ}φµ0 µ00 (5.64)
176 Group Theory in Particle, Nuclear, and Hadron Physics

Therefore these CFP are independent of m and n. This is a basic connec-


tion between the CFP and the ISF and so between the groups SU (n) and
S(f ).
Next we transform the Yamanouchi basis of S(f ) in to the S(f ) ⊃ S(f1 ) ×
SU (f2 ) basis by

0 00
{ν}
 X  {ν}  
{ν}, τ {ν0 } {ν }


m, β{σ}X1 {µ}X2 = 00
00 00
m m m
ν m τ

{ν}
0 0 00 00 (5.65)
τ {ν }m {ν }m , β{σ}X1 {µ}X2

Some lengthy algebra gives the SU (mn) ⊃ SU (m) × SU (n) CFP as,
{ν}τ,β{σ}θ{µ}φ 0 P {ν}β,m {ν 0 }β 0 ,m0 {ν 00 }β 00 ,m00
C{ν 0 }β 0 σ0 µ0 ,{ν 00 }β 00 σ00 µ00 = fmix.m
P
m1 m2 m00 00 Cσm1 ,µm2 Cσ0 m0 ,µ0 m0 Cσ00 m00 ,µ00 m00
1 m2 1 2 1 2
0 00 0 00
  
{ν} τ {ν } {ν } {σ} {σ}, θ{σ0 } {σ 00}

{ν},
m m0 m00 m1 m1 m1
φ{µ0 } {µ00 }
 
{µ}
{µ},
m2 m02 m00
2

where the C on the RHS are the Clebsch-Gordan Coefficients.


We shall use this CFP and ISF connection in determining the hidden colour
components of the wave function in the multiquark systems.
This complex looking expression simplifies for the case of totally antisym-
0 00
metric irreducible representations {ν } = {1f1 }, {ν } = {1f2 }, {ν} = {1f }
whence,

  12
{1f },{σ}θ{µ}φ hσ0 hσ00
C{1f1 }{σ0 }{µ0 },{1f2 }{σ00 }{µ00 } = δσ̃µ δσ̃0 µ0 δσ̃00 µ00 δθσ (5.66)

where the hσ0 , hσ00 and hσ , are the dimensions of the irreducible represen-
tations {σ 0 },{σ 00 } and {σ} of the permutation groups S(f1 ),S(f2 ) and S(f ),
respectively.

5.9 Appendix D: Representation Theory


In quantum mechanics the state vectors are orthonormal vectors |φ1 >
. . . |φn > forming a complete set in vector space Ln . So a state vector ψ > is
[25] , [23]
X
|ψ >= ai |φi > (5.67)
i
Young Diagram 177

The scalar product of basis vector |φi > and |φj > giving orthonormality
condition

Z
< φi |φj > = φ∗i (x)φj (x)d~x
(5.68)
=< φj |φi >= δij
X
|φi >< φi | = 1 (5.69)
i

Thus X X
|ψ >= |φi >< φi |ψ >= ai |φi >
i i

and Hermitian conjugate


X X
< ψ| = < ψ|φi >< φi | = a∗i < φi | (5.70)
i i

Such a space L on which a scalar product as defined is called a unitary


space. We use here only unitary spaces and orthonormal basis vectors.
Let a linear operator A act on vector ~x in space L,

~y = A~x (5.71)
Being linear

A(~x + ~z) = A~x + A~z


(5.72)
A(αx) = αA~x

Note that the above is independent of a coordinate system and is an in-


trinsic definition. So in a “pure” method we may, with a linear vector space
in a basis-independent manner, extract the intrinsic properties in a “pure”
mathematical manner. Nature however, provides “coordinate systems” and
physicists have learned from experience that it (nature) is willing to reveal its
secrets as long as we follow its diktats. History of physics is replete with failed
attempts by physicists who tried to impose their own “beautiful mathemati-
cal structures ” on nature. Pure mathematical structures are powerful, good
and useful guides, but the ultimate arbiters in group theory turn out to be a
proper and judicious application of the representation theory. A major thrust
of the application of group theory in physics is to extract and dig out those
representations which map nature as closely as possible.
Let a quantum mechanics space be defined as above. Let the space be
invariant as well and therefore A|φi > is given as a linear combination,
X
A|φi >= Dji (A)|φj > (5.73)
j
178 Group Theory in Particle, Nuclear, and Hadron Physics

where due to orthonormality of the basis vectors

Dji (A) =< φj |A|φi >, (i = 1, 2, . . . n) (5.74)

The matrix D(A) consisting of elements like Dji (A) is a representation


of the operator A in the representation {φi }. So A acting on a state vector
|ψ >,

X
A|ψ > = A ai |φi >
i
X
= ai A|φi > (5.75)
i
X
= Dji (A)ai |φj >
i,j

Next evaluate AA−1 ,


X
< φi |A|φj >< φj |A−1 |φi > = δik (5.76)
j

Thus
D(A−1 ) = (D(A))−1 (5.77)
−1
So the representation of the inverse operator A is equal to the inverse
of the representation of A. Next
P
< φi |AB|φk >= j < φi |A|φj >< φj |B|φk >

we find

D(AB) = D(A)D(B) (5.78)


For an operator A in L, we define A† as an adjoint or Hermitian conjugate
operator of A as

< φj |A† |φi >=< Aφj |φi > (5.79)


Then we obtain,
∼∗
D(A† ) = D (A) = D† (A) (5.80)

i.e., the representation of the adjoint operator A is the Hermitian conjugate
of the representation of A. For the Hermitian or self-adjoint operator as A† =
A−1

< φj |A|φi > = < Aφj |φ > (5.81)


Young Diagram 179

Thus in an orthonormal basis, a Hermitian operator representation is a


Hermitian matrix

D(A) = D† (A)
A is a unitary operator if

< Aφ|Aφ0 > = < φ|φ0 > (5.82)


Thus,

< Aφ|Aφ0 > = < φ|A+ A|φ0 > = < φ|φ0 >
Now using the definition of a unitary operator (see Appendix B)

A† = A−1 , A† A = AA† = I

D(A† ) = D(A−1 ) and D† (A) = D−1 (A) (5.83)


That is, the representation of a unitary operator is a unitary matrix.
On a change of basis

X
|φ0i > = bij |φj >
j
X
T hen A|φ0i > = bij A|φj >
j
X
= bij Dkj (A)|φk >
jk
X
= bij Dkj (A)(B −1 )kl |φ0l >
jkl
X
= D0 (A)li |φ0l >
l

So the representation of A transforms as


∼ −1 ∼
D0 (A) = B D(A)B (5.84)
† −1 ∗
and for a unitary transformation B = B we get (with C = B )

D0 (A) = CD(A)C −1 (5.85)


Next, for a group G, a representation is a homomorphism of G onto a
group D(G). The elements of D(G) are operations in a space L where the basis
defines the representation space, thereby yielding the matrix representation
D(G). D(G) is linear for linear operators.
180 Group Theory in Particle, Nuclear, and Hadron Physics

As a homomorphic mapping

D(AB) = D(A)D(B) , A, B ∈ G
D(e) = I (the unit matrix)

whence,

D(A−1 ) = (D(A))−1 (5.86)


A representation is denominated faithful if the mapping G → D(G) is an
isomorphism. If D(A) = 1 for all A ∈ G it is termed an identity repre-
sentation. A group has trivially an identity representation.
D∗ (G) is called conjugate representation. If D = D∗ then D is a real
representation. If D∗ is equivalent to D then D is a self-conjugate represen-
tation.
∼ −1
D (G) is called contragradient representation. For unitary represen-
∼ −1
tation D (G) = D∗ (G) and so the contragradient representation is actually
the conjugate representation.
Now if the matrices of a representation are unitary then it is denominated
a unitary representation.

D† (Aα) = D−1 (Aα), α = 1, 2, ....g (5.87)


Two representation D(G) and D (G) of group G in the space L and L0 are
0

equivalent if they obey

D0 (A) = R D(A) R−1 R∈G (5.88)


Note that with a square matrix R, all the equivalent representations have
the same dimension. Note that if the spaces L and L0 are identical then Equa-
tion 5.88 is just a change of basis.
Now trace of a D(A) is called character of element A in the representa-
tion D, and given as

χ(A) = T rD(A) (5.89)


Note that trace is an invariant of the similarity transformation

X X
−1
T rD(A)0 = 0
Dii (A) = Rik Dkl (A) Rli
i i,k,l
X X
= δkl Dkl (A) = Dll (A) (5.90)
k,l l

= T rD(A)

Note that χ(e) is equal to the dimension of the representation.


Young Diagram 181

We can build higher dimensional representation from lower dimensional


ones. Let D1 (G) and D2 (G) of dimensions n1 and n2 , respectively, be two
representations of G. A representation D of dimension n1 + n2 is
 1 
D (A) | 0
D(A) =  − − − | − − −  (5.91)
0 | D2 (A)
Note that the new D(A) actually forms a representation of G by multiply-
ing the block-diagonal matrices,
 1
D (A)D1 (B) |

0

 | 

D(A)DB) =  − − − − − | − − − − − 

 (5.92)
 | 
0 | D2 (A)D2 (B)
D is a direct sum of D1 and D1 ,

D = D1 ⊕ D2 (5.93)
Clearly the character is

χ(A) = χ1 (A) + χ2 (A) (5.94)


The converse of the alone gives an important property of the representation
theory. If for a particular D(G), we may find similarity transformation S
which brings all the matrices D(G) to the same block-diagonal form, then the
representation is called reducible. If it is not possible to find such a similarity
transformation then the matrix D is termed irreducible.
Thus given

D1 (Rα ) |
 
0
S D(Rα ) S −1 =  − − − | − − −  , α = 1, 2, ..g (5.95)
0 | D2 (Rα )

Then D(G) is said to be reduced to a direct sum of two representations


D1 and D2 , i.e., D = D1 ⊕ D2 . D(G) in the block-diagonal form is said to be
in a reduced form.
Another definition – if L is a representation space, and we have two sub-
space L1 and L1 invariant under group D, then L is said to be reducible into
a direct sum of two spaces L = L1 ⊕ L2 , otherwise the space L is irreducible.
In an irreducible basis |νm > (ν gives representation and m = 1, 2, ...nν is
the component index ), then the irreducible matrices are given as
ν
Dm,m0 (A) = < νm |A|νm0 > (5.96)
If for G the only invariant subspaces are the null space and the subspace
182 Group Theory in Particle, Nuclear, and Hadron Physics

itself, then the subspace is called minimal. A minimal invariant subspace of


G is called an irreducible space of G.
Note that each subspace in block diagonal form can be treated indepen-
dently. Thus the transformation

S10
 
| 0 }n1
 − − − | − − −− 
0 | 1 }n2
only changes D1 independently without touching D2 . Choose S10 such that
D(1) becomes block-diagonal. We continue to reduce until D(G) becomes a
direct sum of only irreducible representations

Dν1 (A)
 

D(A) =  Dν2 (A) 


.
X
(νi )
(5.97)
or D(A) = ⊕D
i
(ν1 )
=D ⊕ D(ν2 ) +, ...
P
Note n = i |νi | where n is dimension of D(n) and |νi | of irreducible
representation νi .
Thus the reducible space L is decomposed into a direct sum of all the
irreducible spaces
X
L = Lν1 ⊕ Lν2 + .... = ⊕Lνi (5.98)
i

Suppose αν is the multiplicity of the representation ν1 then


X
D= ⊕αν D(ν)
ν

Thus the character


(ν)
X
χi = αν χi (5.99)
ν

where for the irreducible representation it is termed a simple or primitive


character.
Note that the problem of determining all the representations of a finite
group or a compact Lie group reduces to the problem of finding all the inde-
pendent irreducible representations of it.

————————————————–
Problem D.1: We know that in a 2-dimension R2 space rotation matrix
is  
cosθ −sinθ
D(θ) = on basis e1 , e2
sinθ cosθ
Young Diagram 183

Is it reducible in R2 ? Under what conditions is it reducible?


————————————————–

Schur’s Lemma: In the definition of equivalence of representation in


Equation 5.88 nothing was said of reducibility or irreducibility of D(A) or
D0 (A). Actually it is easy to see that two equivalent representations must
be both reducible or both irreducible. For irreducible representations, Schur’s
Lemma provides a useful test of equivalence. It shows that if D(A) and D1 (A)
are irreducible representations of a group G or linear spaces L and L0 , and if
there is a linear map S:L → L0 such that

D0 (g)S = SD(g) f or all gG (5.100)


then either
(a) D(A) and D0 (A) are inequivalent and S=0; or
(b) D(A) and D1 (A) are equivalent , then S 6= 0 and S −1 exists.
Part (a) is intuitively obvious. If D(A) and D0 (A) are nonequivalent, then
any S that one may obtain should be zero. If we know that S is non-zero, then
in fact S should be non-singular, and the representation must be equivalent.
Clearly the operator S should not be non-zero and singular at the same time
if both the representations are irreducible.
Take the special case of D0 (A) = D(A), so we have just one irreducible
representation on a given space L. Then clearly for D(A) irreducible

D(g)S = SD(g) f or all g ∈ G (5.101)


means S = λI, i.e., multiple of a unit operator in L. This is again Schur’s
Lemma. Thus if a matrix commutes with all the matrices of an irreducible
representation, then the matrix is necessarily a multiple of the unit matrix.
Schur’s Lemma is basic in group representation theory, as it gives necessary
and sufficient conditions for irreducibility. Thus a representation is irreducible
if and only if the only matrices, which commute with all the representation,
are multiples of the unit matrix. Note that the converse of Schur’s Lemma is
not correct.
Direct product of two irreducible representations of group G. Let two
representations D1 (G) and D2 (G) operate on vector spaces L1 and L1 , re-
spectively. They are both either complex or both real. The direct product
vector space L1 × L2 is complex or real depending upon the nature of L1 and
L2 . Given x1 ∈ L1 and x2 ∈ L2 , L1 × L2 is spanned by outer products of
ordered pairs of vectors of the form x1 ⊗ x2 . One defines the direct product
representation
D(G) = D1 (G) ⊗ D2 (G) as

D(g)(x1 ⊗ x2 ) = D1 (g)x1 ⊗ D2 (g)x2 (5.102)


Similarly for the whole of L1 × L2 by the property of linearity.
184 Group Theory in Particle, Nuclear, and Hadron Physics

The irreducibility of D1 (G) and D2 (g) does not guarantee irreducibility


of D(G). In fact on general grounds D(G) may be reducible. If it is fully re-
ducible, then various irreducible representation may occur with various mul-
tiplicities. This is given as Clebsch-Gordan series and is the direct sum of
irreducible representations of the product D1 (G) × D2 (G).

————————————————–
Problem D.2: Show that the irreducible representations of an Abelian
group are all one-dimensional.
————————————————–

5.10 Appendix E: Symmetry in Quantum Mechanics


If a system is described by a wave function ψ(x, t) and given the Hamilto-
nian H, the Schroedinger equation is [32], [20], [44]


Hψ = i~ ψ (5.103)
∂t
Its complex conjugate Schroedinger equation is
∂ ∗
Hψ ∗ = −i~ ψ (5.104)
∂t
The expectation value at time t of an observable θ is
Z
< θ(t) >= ψ ∗ (t)θψ(t)d3 x (5.105)

Next, the rate of change of θ is, due to Equation 5.103 and 5.104,
Z

i~ θ(t) = ψ ∗ (t)(θH − Hθ)ψ(t)d3 x (5.106)
∂t
Define the commutator of θ and H

[θ, H] = θH − Hθ (5.107)
∂θ
This is the expression in the bracket on the right-hand side above. If ∂t =0
then θ is conserved and then,

[θ, H] = 0 (5.108)
Then the operator θ is said to commute with the Hamiltonian H. Thus the
observable associated with an operator θ is conserved if θ commutes with the
Hamiltonian of the system.
Young Diagram 185

Consider an infinitesimal displacement of δx in x direction of a one-


dimensional system

x → x = x − ∆x => x = x0 + ∆x (5.109)
the wave function becomes

ψ 0 (x0 ) = ψ(x) = ψ(x0 + ∆x)


as x0 is a dummy variable

ψ 0 (x) = ψ(x + ∆x)

By Taylor expansion and retaining only the first order term in δx

∂ i
ψ(x + ∆x) ≈ ψ(x) + ∆x ψ(x) = (1 + ∆x px ) ψ(x) (5.110)
∂x ~
where
~ ∂
px = (5.111)
i ∂x
which we recognize as the momentum operator in quantum mechanics. It is
called the generator of a translation along the x-axis.

————————————————–
Problem E.1: Demonstrate that retention of all the terms in the above
Taylor expansion leads to ψ 0 (x) = U ψ(x) where U is a unitary transformation

U = exp(i∆x px /~)

————————————————–

If now the system is invariant under an infinitesimal ∆x, x → x + ∆x,


then this means that the expectation value of the Hamiltonian H shall remain
unchanged
Z Z
ψ (x)H ψ(x)d x = ψ ∗ (x + ∆x)H ψ(x + ∆x)d3 x
∗ 3

Z
i i
ψ ∗ (x)(1 − ∆x px )H(1 + ∆x px )ψ(x)d3 x
~ ~
This is only possible if
186 Group Theory in Particle, Nuclear, and Hadron Physics

i i
H = (1 − ∆x px ) H (1 + ∆x px )
~ ~ (5.112)
i
= H − ∆x [px , H] + O(∆x )2
~
Ignoring terms of order (∆x)2 , then the above is possible if

[px , H] = 0

Thus it means that px is conserved.


Thus we have proven that invariance under infinitesimal space translation
(along the x-axis) implies conservation of the corresponding momentum. We
can similarly demonstrate conservation of momentum for translation along y-
and z-directions as well. Thus invariance under infinitesimal translation along
any arbitrary direction leads to the corresponding momentum conservation.
We saw in Chapter 1 that as shown here conservation of linear momentum
means that the space is symmetric under translation and this arises because we
cannot measure the absolute origin of space. Thus unmeasurability of absolute
position in space leads to homogeneity of space, i.e., translational invariance
and thereby leading to momentum conservation.

————————————————–
Unsolved Problem 5.2: As performed for spatial translations, for a time
translation, determine that its generator is the Hamiltonian operator H =

i~ ∂t . Demonstrate that invariance under time translation gives conservation
of energy.
————————————————–

Next consider a rotation around the z-axis. We take from Chapter 3 for
SO(2) group,
 0    
x cosθ sinθ x
=
y0 −sinθ cosθ y
Consider infinitesimal rotations

x0 = x + y∆φ = x + ∆x
y 0 = y − x∆φ = y + ∆y
z0 = z

The wave function goes as


Young Diagram 187

~
ψ(r0 ) → ψ(r) + ∆~r.∇ψ(r 0
) + ...
∂ψ(r) ∂ψ(r) ∂ψ(r)
= ψ + ∆x + ∆y + ∆z
∂x ∂y ∂z
∂ ∂
= [1 + (y − x )∆φ]ψ(r) + . . .
∂x ∂y
We know that the orbital angular momentum operator
~ ∂ ∂
Lz = (~r × p~)z = xpy − ypx = (x −y )
i ∂y ∂x
so
i
ψ(r) → [1 − Lz ∆φ]ψ (5.113)
~
So we can identify the generator of rotations as the angular momentum
operator.
Just as above, we can show that invariance under rotations about the z-axis
implies conservation of Lz

[Lz , H] = 0 → Lz conserved.
A rotation through a finite angle φ can be considered as a product of N
rotation through φ/N . Then for finite rotations,

ψφ (x) ∼ (1 − Lz ) ψ(x, y, z)
N
In the limit N → ∞, the above becomes

ψφ (r) = R(φ)ψ(r)
with

R(φ) = e−iφLz (5.114)


The same could have been found by integrating the above equation.
Thus the effect of rotation about the z-direction is obtained by exponen-
tiation of the angular momentum operator in that direction. We generalize to
the case where φ ~ indicates the axis of rotation and its magnitude as

~ = e−iφ.J~ ~
R(φ) (5.115)
where we generalize from L to J (in anticipation of spin degree of freedom).
This Lie group, labelled SO(3) and called the special unitary group in three
dimensions. In quantum mechanics this group is well specified by its generators
which we saw are the angular momentum operators, given by Ji , i = 1, 2, 3
with Lie algebra
188 Group Theory in Particle, Nuclear, and Hadron Physics

[Ji , Jj ] = iijk Jk (5.116)


A Casimir operator of SO(3) is

J~2 = J12 + J22 + J3 (5.117)


which commutes with all the generators of SO(3)

[J~2 , Ji ] = 0, i = 1, 2, 3 (5.118)

————————————————–
Problem E.2: Show that [J~2 , Ji ] = 0, i=1,2,3
————————————————–

Thus as J~2 and Jz can be diagonalized for a quantum mechanical state


given by |j, m >,

J~2 |f, m > = j(j + 1)|j, m >


Jz |j, m > = m|jm >

As J~2 is Hermitian both (j, m) are real and eigenstate corresponding to


different (j, m) values are orthogonal.

————————————————–
Problem E.3: By defining the raising and lowering operators J± = J1 ±
iJ2 and using the above commutation rules, prove that (j, m) are both integers
or half integers. Explain why we get here spin 21 when we are dealing with
SO(3).
————————————————–

Thus we are able to provide a complete specification of the properties


of the angular momentum operator J~ for a given value of j. We do so by
giving matrix elements of J~ as < j, m | Ji | j, m >. In general we define
a matrix representation of dimensionality n as a set of three nxn matrices
of Ji satisfying the commutation relation Equation 5.117. Given a particular
matrix representation we can obtain an infinite number of others by similarity
transformation Ji0 = U Ki U † where U is a unitary matrix. From among all
these equivalent representations, we but need just one to study the state.
We use an appropriate U to obtain a representation with Jz diagonal (as
Jz is Hermitian). The basis vectors of this diagonal matrix representation will
be eigenvector of Jz and J~2 and labelled as |j, m > .
It we have a representation in which J~2 possesses more than a single eigen-
value, then it shall be reducible. By appropriate choice this may be reduced
Young Diagram 189

to provide irreducible representation in terms of Ji placed in block diagonal


form
 
(j1 ) | |
 −−− | −−− | −− 
 
Ji =  | (j2 ) | 

 −− | −−− | −−− 
|

Note that the reduction of a reducible representation to block diagonal


form is exactly what is required in the addition of angular momenta. So say
we have to combine two sets of states |j1 , m1 > and |j2 , m2 >. Then the
various j values and given by |j1 − j2 | 6 j 6 j1 + j2 . The transformation
which leads to the proper reduction defines the Clebsch-Gordan coefficients
< j1 , m1 j2 m2 |jm >
X
|jm >= < j1 m1 j2 m2 |jm > |j1 m1 > |j2 m2 >
m1 m2

————————————————–
Problem E.4: Quantum nonlocality, quantum jumps, wave-function col-
lapse; essentially all puzzling aspects of quantum mechanics, were hypothe-
sized by Heisenberg to occur in a ”potentia” space. Show that parallel to the
real three dimensional SO(3)l space there is a coexisting dual space called
potentia space SO(3)p , wherein velocity c → ∞. Show that gravity actually
sits in the space of potentia. The space of potentia does not allow gauging.
Thus gravity is not quantized.
————————————————–

5.11 Solutions of Problems

Solution 5.1:
n = 4 partitions as [4], [31], [22], [211], [1111]. So YD are

[4] ≡ , [31] ≡ , [22] ≡ ,


190 Group Theory in Particle, Nuclear, and Hadron Physics

[211] ≡ , [1111] ≡
————————————————–
Solution 5.2:
[4] ≡ 1 2 3 4 ; 1-dimensional representation

1 2 3 1 3 4 1 2 4
[31] ≡ 4 ; 2 ; 3 ; 3-dimensional representation

1 2 1 3
[22] ≡ 3 4 ; 2 4 ; 2-dimensional representation

1 2 1 3 1 4
3 2 2
[211] ≡ 4 ; 3 ; 3 ; 3-dimensional representation

1
2
3
[1111] ≡ 4 ; 1-dimensional representation

(Check 12 + 32 + 22 + 32 + 1 = 1 + 9 + 4 + 9 + 1 = 24 = 4!)

————————————————–
Solution-5.3:
n = 5 partitions:

[5] ≡ ; [41] ≡ ;

[32] ≡ ; [311] ≡

[221] ≡ ; [2111] ≡ ; [1111] ≡


Young Diagram 191

So seven irreducible representations,

[5] ≡ 1 2 3 4 5 ; 1 − dim.rep.

1 2 3 4 1 3 4 5 1 2 4 5 1 2 3 5
[41] ≡ 5 ; 2 ; 3 ; 4 ;
4 − dim.rep.

1 2 3 1 3 4 1 2 4 1 2 5 1 3 5
[32] ≡ 4 5 ; 2 5 ; 3 5 ; 3 4 ; 2 4 ;
5 − dim.rep.

1 4 5 1 2 5 1 3 5 1 3 4 1 2 4 1 2 3
2 3 2 2 3 4
[311] ≡ 3 ; 4 ; 4 ; 5 ; 5 ; 5 ;
6 − dim.rep.

1 4 1 3 1 3 1 2 1 2
2 5 2 5 2 4 3 5 3 4
[221] ≡ 3 ; 4 ; 5 ; 4 ; 5 ; 5 − dim.rep.

1 5 1 4 1 3 1 2
2 2 2 3
3 3 4 4
[2111] ≡ 4 ; 5 ; 5 ; 5 ; 4 − dim.rep.

1
2
3
4
[1111] ≡ 5 ; 1 − dim.rep.

(Check 12 + 42 + 52 + 62 + 52 + 42 + 12 = 1 + 16 + 25 + 36 + 25 + 16 + 1+ =
120 = 5!)

————————————————–
Solution 5.4:

S6 = [6], [51], [42], [411], [33], [321], [3111], [222], [2211], [21111], [111111]

Given YD’s as follows with the corresponding dimensions:


192 Group Theory in Particle, Nuclear, and Hadron Physics

6.5.4.3.2.1
[6] ≡ ; d= 1.2.3.4.5.6 =1

6.5.4.3.2.1
[51] ≡ ; d= 1.2.3.4.6.1 =5

6.5.4.3.2.1
[42] ≡ ; d= 1.2.4.5.1.2 =9

6.5.4.3.2.1
[411] ≡ ; d= 1.2.3.6.2.1 = 10

6.5.4.3.2.1
[33] ≡ ; d= 2.3.4.1.2.3 =5

6.5.4.3.2.1
[321] ≡ ; d= 1.3.5.1.3.1 = 16

6.5.4.3.2.1
[3111] ≡ ; d= 1.2.6.3.2.1 = 10

6.5.4.3.2.1
[222] ≡ ; d= 3.4.2.3.1.2 =5

6.5.4.3.2.1
[2211] ≡ ; d= 2.5.1.4.2.1 =9

6.5.4.3.2.1
[21111] ≡ ; d= 1.6.4.3.2 =5
Young Diagram 193

6.5.4.3.2.1
[111111] ≡ ; d= 6.5.4.3.2.1 =1

(Check d = 12 + 52 + 92 + 102 + 52 + 162 + 102 + 52 + 92 + 52 + 12

= 1 + 25 + +81 + 100 + 25 + 256 + 100 + 25 + 81 + 25 + 1 = 720 = 6!)

————————————————–
Solution 5.5:
P
ψa = Ya (abcd) = [1 − (14)][ P (123)](abcd)

= {a(1)b(2)c(3)+a(1)b(3)c(2)+a(2)b(1)c(3)+a(2)b(3)c(1)+a(3)b(1)c(2)+
a(3)b(2)c(1)}d(4)

−{a(4)b(2)c(3)+a(4)b(3)c(2)+a(2)b(4)c(3)+a(2)b(3)c(4)+a(3)b(4)c(2)+
a(3)b(2)c(4)}d(1)

= abcd + acbd + bacd + cabd + bcad + cbad − dbca − dcba − dacb − dabc −
dcab − dbac
P
ψb = Yb (abcd) = [1 − (13)][ P (124)]abcd

= abcd + adcb + bacd + dacb + bdca + dbca − cdab − cbad − cabd − cadb −
cdba − cbda
P
ψc = Yc (abcd) = [1 − (12)][ P (134)]abcd

= abcd + abdc + cbad + dbac + cbda + dbca − bacd − bdca − abdc − bcda −
bdac − bcad

ψd = Yd (abcd) = [1 − (24)][1 − (13)][1 + (34)][1 + (12)]abcd

ψe = Ye (abcd) = [1 − (34)][1 − (12)][1 + (24)][1 + (13)]abcd

Proper orthogonalization would be required for the last two cases.


————————————————–
Solution 5.6:
194 Group Theory in Particle, Nuclear, and Hadron Physics

⊗ ⊗ = ⊕ ⊕

3.4.5
≡ d= 1.2.3 = 10

3.4.2
≡ d= 1.3.1 =8

3.2.1
≡ d= 3.2.1 =1

So, 3 ⊗ 3 ⊗ 3 = 10 ⊕ 8 ⊕ 8 ⊕ 1

————————————————–
Solution 5.7:
For SU (n) a single box stands for fundamental n-dimensional representa-
tion. Its conjugate n̄ is given by a single column with (n − 1) number of boxes.
The product of these gives a singlet with a saturated column of n-boxes and
another with first row with two boxes and (n − 1) boxes in the first column.
We calculate the latter YD dimensions as
n(n+1)(n−1).....2.1
d= 1.n.(n−2)....2.1 = (n + 1)(n − 1) = n2 − 1

Finally, n̄ × n = 1 ⊕ (n2 − 1)

————————————————-
Solution 5.8:
1 1
1 1 1 1 1
(a). ⊗ 1 1 1 = ⊕ 1 ⊕ 1 ⊕
1
1
1

⊗ = ⊕ ⊕ ⊕

8 ⊗ 10 = 35 ⊕ 27 ⊕ 10 ⊕ 8

1 1
1 1 1 1
(b). ⊗ 1 1 1 = ⊕ ⊕
Young Diagram 195

1 1 ⊕ 1 1 1

⊗ = ⊕ ⊕ ⊕•

¯ ⊗ 10 = 64 ⊕ 27 ⊕ 8 ⊕ 1
10

————————————————-
Solution 5.9:
We have 4 occurring in the above YT as follows:

4 4 4
4 4
⊕ ⊕ ⊕ 4 ⊕ ⊕ 4 ⊕
4
4 4
4 ⊕ 4

For SU(4)→ SU (3) reduction drop boxes with 4 in it.

⊕ ⊕ ⊕ ⊕ ⊕ ⊕ ⊕

4.5.6.3.4.2
in SU (4) d = 3.5.3 = 64

This reduces to SU (3) diagrams,

⊕ ⊕ ⊕ ⊕ ⊕ ⊕ ⊕

So finally,

¯ +8
64 ⊃ 8 + 3̄ + 6 + 15 + 3 + 6̄ + 15

————————————————-
Solution 5.10:
196 Group Theory in Particle, Nuclear, and Hadron Physics

→ ⊕ 3 ⊕ 3 3 ⊕ 3 3 3

⊃ ⊕ ⊕ ⊕•

10 ⊃ 4 + 3 + 2 + 1

Now ‘3’ in SU (3)F quark model we identify as 1 = u, 2 = d, 3 = s-


quarks. So these diagrams tell us the SU(2) structure with how many s-quarks
(i.e., 3) are there. So the SU (3)f decuptet of the quark model has: a 4-plet
spin 32 of strangeness free baryons ∆++ , ∆+ , ∆0 , ∆− ; strangeness-one triplet
? ? ? ? ?
Σ+ , Σ0 , Σ− ; strangeness-two doublet Ξ− , Ξ0 ; and a strangeness-three Ω− ,
which is an isospin singlet.
————————————————-
Solution 5.11:

→( , •) ⊕ ( , )⊕( , ) ⊕ (• , )

56 → (20, 1) ⊕ (10, 2) ⊕ (4, 3) ⊕ (1, 4)

in SU(6) and SU(4), d=56 and 20, respectively.

in SU(4), d=10.

And,

→( , •) ⊕ ( , )⊕( , ) (5.119)
20 → (4̄, 1) ⊕ (6, 2) ⊕ (4, 1)

in SU(6) has dimension 20.

————————————————–
Solution 5.12:

→( , •) ⊕ ( , )⊕( , ) ⊕ (• , )

56 → (10, 1) ⊕ (6, 3) ⊕ (3, 6) ⊕ (1, 10)


Young Diagram 197

→( , •) ⊕ ( , )⊕( , ) ⊕ (• , ) (5.120)
20 → (1, 1) ⊕ (3̄, 3) ⊕ (3, 3̄) ⊕ (1, 1)

————————————————–
Solution 5.13:
Under SU (8)SF ⊃ SU (4)F ⊗ SU (2)S

→( , )
8 → (4, 2)

Next two particle states,

⊗ = ⊕
8 ⊗ 8 = 36 ⊕ 28

Under SU (8)SF ⊃ SU (4)F ⊗ SU (2)S reduction,

→( , )⊕( , )

36 → (10, 3) ⊕ (10, 1)

→( , )⊕( , )

28 → (10, 1) ⊕ (6, 3)

Baryons made up of three fundamental representations of SU (8) are

⊗ ⊗ = ⊕ ⊕ ⊕

8 ⊗ 8 ⊗ 8 = 120 ⊕ 168 ⊕ 168 ⊕ 56


198 Group Theory in Particle, Nuclear, and Hadron Physics

These are reduced as follows

→( , )⊕( , )

120 → (20, 4) ⊕ (20, 2)

→( , )⊕( , )

56 → (4̄, 4) ⊕ (20, 2)

→( , )⊕( , )⊕( , )⊕

( , )

168 → (20, 4) ⊕ (20, 2) ⊕ (20, 2) ⊕ (4̄, 2)

————————————————–
Solution 5.14:
In SU (6)SF model,

⊗ = ⊕
6 ⊗ 6 = 21 ⊕ 15

S-wave symmetric state is 21-dimensional irreducible representation. In


terms of SU (6)SF ⊃ SU (3)F ⊗ SU (2)S reduction

=( , )⊕( , )
21 = (6, 3) ⊕ (3̄, 1)

Now let the third quark attach to the symmetric diquark 21-dimensional
state to form three quark baryons.
Young Diagram 199

⊗ = ⊕
21 ⊗ 6 = 56 ⊕ 70

Next we attach SO(3)l orbital angular momentum state (see Chapter 7).
Then the symmetric states belonging only to the multiplets (56, L+ even ) and
(70, L−even ) would be observed. In earlier days of quark model (1970s) the bary-
onic resonances did appear to obey this constraint. But this is now overruled
by the full quark model which has many more excitations than just these two
blocks (see Chapters 7 and 9).
————————————————–
Solution D.1:
D(θ) is defining one-parameter representation of the group SO(2) on R2 .
As D(θ)I = ID(θ) and as per Schur’s Lemma it is irreducible on R2 . But
there exists a fully reducible representation of SO(2) on complexification of
R2 , i.e., C2 treated as a vector space. As (e1 , e2 ) is a basis in R2 , it is also a
basis on C2 . Let us change to basis:

1
φ01 = − √ (e1 + ie2 ) = rY1,1 (θ, φ)
2
(5.121)
1
φ02 = √ (e1 − ie2 ) = rY1,−1 (θ, φ)
2
The transformation now is
!
− √12 − √i2
B=
+ √12 − √i2

Then in the new basis


e−iθ
 
∼ 0
D0 (θ) = B ∗ D(θ)B =
o eiθ

Thus the two-dimensional representation is reduced to two one-dimensional


representation in C2 space.
————————————————-
Solution D.2:
For a fixed g ∈ G, the Abelian property implies that

D(g)D(g 0 ) = D(g 0 )D(g) f or all g 0 ∈ G

Here D(g 0 ) plays the role of S in Equation 5.100 of Schur’s Lemma. Then it
implies that
D(g 0 ) = λI
200 Group Theory in Particle, Nuclear, and Hadron Physics

As g 0 is an arbitrary element of G, D(g) is diagonal for all g ∈ G. Since


D is irreducible, the matrix D(g) must be |x|, i.e., the representations are
one-dimensional.
————————————————–
Solution E.1:

ψ 0 (x) = ψ(x + ∆x)


∂ 1 ∂2
= ψ(x) + ∆x ψ(x) + ∆x2 2 ψ(x)
∂x 2 ∂x
3
1 ∂
+ ∆x3 3 ψ(x) + . . .
6 ∂x

= exp(∆x )ψ(x)
∂x
so ψ 0 (x) = U ψ(x)

where U = exp(i∆x px /~)


————————————————–
Solution E.2:
[J~2 , Ji ] = [Jj Jj , Ji ] = iJj jik Jk + ijik Jk Jj

= ijik (Jj Jk + Jk Jl )

As ijk is antisymmetric and the expression (Jj Jk + Jk Jj ) is symmetric,


we obtain zero on the right-hand side.
————————————————–
Solution E.3:
J~2 |j, m >= j(j + 1)|j, m >
Jz |j, m >= m|j, m >
J± = J1 ± J2 and [Jz , J± ] = ±J±
(Jz J± − J± Jz )|j, m >= ±J± |j, m >
Jz {J ± |j, m >} = (m ± |){J ± |j, m >}

Thus J± |j, m > is an eigenstate of J3 with eigenvalue (m ± 1). Hence it


must be some multiple of the state |j, m ± 1 >.
J± |j, m >= a± |j, m ± 1 >

J~2 = J± J± + Jz2 ± Jz

< j, m|J± J± |j, m > = |a± |2 < j, m ± 1|j, m ± 1 >
=< j, m|J~2 − Jz (Jz + 1)|j, m >
2
|a± | = j(j + 1) − m(m ± 1)

Choosing a positive
p sign,
J± |j, m >= j(j + 1) − m(m ± 1)|j, m ± 1 >
Young Diagram 201

Demanding j(j + 1) − m(m ± 1) > 0 leads to −j 0 m 0 j


This last condition demands that there is a state with maximum m+ in
which J+ |j,
pm+ >= 0 and similarly a minimum m− with J− |j, m− >= 0. In
that case j(j + 1) − m± (m± ± 1) = 0, so m± = ±j. Now m+ and m− can
clearly be reached by application of J± on |j, m > a finite number of times,
that is an integral number of steps. So (m+ − m− ) = 2j must be an integer.
So m takes on values (−j, −j + 1, ...j − 1, j) for j integer or half integer.
The question is how do we get here spin 1/2, etc. while the group was still
that of SO(3)? The reason is that as we shall see, the Lie algebra associated
with SO(3) is isomorphic to the Lie algebra of SU (2). So as far as the algebra
goes, the same algebra works for both the integral and the half-integral spins.
This is the so-called double-valued representation of the SO(3) Lie group.
The existence of double-valued representation of SO(3) is intimately tied to
the double-connectedness of the group manifolds, i.e., how the groups SO(3)
and SU (2) map onto each other globally. This is discussed in Chapter 3 for
SO(3) group. The answer would be related to the fact that the irreducible rep-
resentations for j-integral and j-half integral were derived using Lie algebras,
which provides group structure as reflected in the vicinity of the identity. Thus
there was no control over the global behaviour of the corresponding group.
Therefore there arises a 2-1, homomorphism between the groups SO(3) and
SU (2). The unitary group SU (2) is said to be the covering group of the rota-
tion group SO(3).
Next, how can we exclude spin-half for pure SO(3) group Lie algebra? To
do this, we have to add one more condition to the above analysis. And this
condition is that for SO(3) the orbital angular momentum zero should be
allowed, i.e., l = 0.
————————————————–
Solution E.4:
Let two electrons be residing in space SU (2)S ⊗ SO(3)l . SO(3)l speci-
fies the three-dimensional x-,y-,z-space. The antisymmetric wave function at
the positions of sequential numbers (12) (Equation 5.29) do not exist in the
ordinary SO(3)l space (see vixra.org/abs/1603.0344).
The group structure SU (2) has a centre of Z2 . Then the factor group,

SU (2)S ∼
= SO(3)p (5.122)
Z2
Here given the group structure space in the above equation, there is no
justification in associating the above orthogonal group with the group SO(3)l .
We should treat it as another independent SO(3) group and is thus labelled
with another subscript ”p”, identified with ”potentia”.
As Z2 is a discrete subgroup of the group SU (2)S and hence it should label
its fundamental representation with its Z2 centre elements [0,1] as,
 
↑ (0)
(5.123)
↓ (1)
202 Group Theory in Particle, Nuclear, and Hadron Physics

In the symmetric group S3 the antisymmmteric state is,

(5.124)
But this is zero for electrons as these are representations of the group
SU (2)S . We do so by putting the Z2 labels in the Young Diagram for the
SU(2) fundamental representation as,

0
1 (5.125)
And thus for three particles the relevant non-zero Young Diagram is,

0 0
1 (5.126)
The labels [0, 1] =∼ [1, 2] are now associated with the sequential labels
(12) corresponding to the center Z2 . The centre being a discrete group, the
exchange over this space is a jump between 1 and 2 with infinite speed c → ∞.
The points at which the particles are defined in both the spaces is what makes
these two to sit piggyback on each other. When measuremt is performed in
our SO(3)l space then the wave function collapse occurs and nonlocality is
manifested.
For a single particle, the phase of the wave function eiφ ψ is relevant. This
gives the group U (1). Now given the additive group of the real number R and
the infinite set of integers Z, the factor group is,
R∼
= U (1) ∼ SO(2)p (5.127)
Z
Now SO(2)l is a subgroup of the orbital space SO(3)l . However we identify
the above SO(2)p as an independent and different space which is labelled by
the set Z. We suggest that this potentia space of SO(2)p labels the particle
in that space by the discrete set Z. Let us propose that the spaces SO(2)l
and SO(2)p are simultaneous and dual to each other and sitting piggyback on
each other.
When observation is made in the space SO(2)l then the wave function
collapses in such a manner that in the potentia space jumps in Z from 1 →
2 → 3 → 4 → · · · occur, while it travels continuously and with velocity v ≤ c
in our orbital space. Clearly for a bound state these jumps would correspond
to instantaeous quantum jumps in the potentia space. So quantum jumps do
not occur in real SO(2)l space but in the SO(2)p potentia space with infinite
speed.
Gravity sits in the space of potentia (see /vixra.org/abs/1603.0411) and
thus is not quantized.
————————————————–
Chapter 6
Quantum Chromodynamics

6.1 Why the SU (3)c Group? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203


6.2 QCD Langrangian and Asymptotic Freedom . . . . . . . . . . . . . . . . . . . . 206
6.3 Colour Singlet States and Confinement . . . . . . . . . . . . . . . . . . . . . . . . . . 211
6.4 Pentaquarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
6.5 Large Nc Colour QCD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
6.6 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224

6.1 Why the SU (3)c Group?


Quantum Electrodynamics is built around the Abelian group U (1)em . Its
counterpart for strong interactions is Quantum Chromodynamics built around
the non-Abelian group SU (3)c . Two questions arise: first, why the number 3
for colour and second, why the particular group SU (3)c and not some other
group like SO(3)c ?
As we shall see in Chapter 7, if the degrees of freedom in the SU (3)F
model of quarks are its 3-flavours, its 2-spins (quarks are spin-1/2 fermions)
of SU (2)s and the orbital degrees of freedom in SO(3)l , then what works
best to describe all the baryons is the symmetric wave function. Now, as
baryons are constituted of three quarks, the total wave function should be
antisymmetric. But in the three degrees of freedom (as above) known in the
early 1960s, the wave function was puzzlingly symmetrical. One can actually
build an antisymmetric wave function within the three degrees of freedom
available, but it fails to reproduce the physical observables of baryons. Only
the symmetric quark model wave function works!
To incorporate this new fact, certain scientists were even willing to alto-
gether abandon the Pauli Exclusion principle for quarks. On the other hand,
the remaining scientists held this well-attested law to be more sacrosanct and
thus they invoked an ad hoc new degree of freedom, denominated ‘colour’, to
take account of this perplexing problem. The idea was to have colour produce
a fully antisymmetric component of the wave function, so that when multi-
plied with the above symmetric quark model wave function, the total wave
function would be antisymmetric. And the three colours work. Yet it was,
at the time, at best a theoretical fix. However, just as quarks know of their

203
204 Group Theory in Particle, Nuclear, and Hadron Physics

1 2 3 4 5

S GeV

FIGURE 6.1: Schematic plot of the ratio R which proves the presence of
three colours

flavour degree of freedom, do they also know of the three-valued colour degree
of freedom?
Indeed they do! The three-valued colour is an experimentally confirmed
quantum number. And hence colour, which was initially an ad hoc concept,
later turned out to be an empirically confirmed reality of hadrons. For exam-
ple:
(A) Calculation shows that the following ratio for quarks without colour
is

σ(e+ e− → hadrons) X
R= + − + −
∼ eq 2 (6.1)
σ (e e → µ µ ) q

where R is the cross-section ratio of σ(e+ e− → all of the hadrons) and


σ (e+ e− → µ+ µ− ), i.e., to a muon pair, is equal to the sum of all charges of
quarks which may exist at that energy available to the system (Figure 6.1).
At low energies, where only v, d, s quarks are available, the RHS is 49 +
1 1 2
9 + 9 = 3 . This is without colour degree of freedom. Experimentally R in
this region however it is not 2/3 as predicted above, but actually ∼ 2. Hence
one has to multiply it by the colour factor of 3 to match the experimental
number. Above the c-quark threshold the quarks are u, d, s, c. Thus there is
an extra factor of 4/9. Without colour the ratio would be 10/9, again in
disagreement with the experimental number, and again an extra factor of 3
from the colour degree of freedom fixes the problem. Similarly for R above
the b-quark threshold. Hence
P including the number of colours on the RHS the
correct expression is Nc q eq 2 .
(B). Looking at the decay of the τ -lepton, there exist three distinct possi-
bilities:
Quantum Chromodynamics 205

τ − → e− ν̄e ντ , τ − → µ− ν̄µ ντ , τ − → dūντ


So if there were no colours, the branching ratio

Γ(τ − → e− νe− ντ )
B0 = (6.2)
Γ(τ − → all)
would be 1/3. Meanwhile, if there were 3 colours then the τ − decay to the
quark sector would carry an extra factor of 3, thus making the ratio B0 to be
1/5. The experimental number B0 matches the value of 1/5 nicely, thereby
confirming the three-valuedness of colours for quarks in an extremely elegant
manner.
(C) An oft-quoted support in favour of Nc = 3 lies in the study of π 0 → γγ
decay. This process takes place through an anomaly in divergence of the axial
vector currents [42]. The decay rate is
2 2
2 α mπ0
Γ π 0 → γγ = Nc2 (Q2u − Q2d )

(6.3)
64 π 3 Fπ2
Qu and Qd are the u− and d− quark charges, Nc is the number of colours,
e2
mπ0 is the neutral pion mass, α = 4π and Fπ , the pion decay constant
∼ 91MeV.
The experimental value of the decay rate is ∼ 7.8eV. The quark charges are
canonically taken as Qu = 23 and Qd = − 31 [42]. If there were no colours, i.e.,
Nc = 1, then from the above formulae one obtains the decay rate of 0.84 eV ,
far too low a value. Thus as per Equation 6.3, one is forced to include Nc = 3
and then the fit is good. This is taken as a standard proof of the evidence of
3-colours in particle physics [[42], p161 ].
However in Chapter 10 we demonstrate that in the Standard Model with
group structure SU (3)c ⊗ SU (2)L ⊗ U (1)Y , the electric charge of quarks
intrinsically knows of the colour degree of freedom. It was also shown that
the correct charge does not merely take the static values of 2/3 and -1/3 (i.e.,
independent of any colour), but is given as
1 1
Qu = (1 + )
2 Nc
1 1
Qd = (−1 + ) (6.4)
2 Nc
For NC = 3 this yields the correct empirically determined charges of 2/3
and -1/3, However it is important to realize that there exists a significant
colour component in it. These colour-dependent charges, and not the static
(colour-independent) 2/3, -1/3 charges, are the correct ones to use for SU(Nc )
QCD theory.
Now the full colour dependence in Equation 6.3 is
206 Group Theory in Particle, Nuclear, and Hadron Physics
"  2   2 #2
2 2 2 2 2 1 1 1 1
Nc (Qu − Qd ) = Nc 1+ − −1 + =1
2 Nc 2 N
(6.5)
And hence overall there is no Nc − dependence left in the decay rate of
π 0 → γγ and the subsequent result matches the experiment well. So when
correct colour dependent electric charges (as in Equation 6.4) for quarks are
taken, the decay rate is actually independent of colour degrees of freedom.
Two points should be noted. First, that the conventional way of stating
that π 0 → γγ decay is a proof of Nc = 3 number of colours is wrong. It is
independent of the colour degree of freedom. Second, that for arbitrary Nc the
current charges for u- and d-quarks are as given in Equation 6.4 and only for
Nc = 3 are these equal to 2/3 and -1/3. Hence electric charges are not static
2/3 and -1/3 but intrinsically depend upon colour.
Electromagnetic force is independent of colour (strong) force but the elec-
tric charge knows of the colour degree of freedom. This is a unique unification
arising in the SU (3)c ⊗ SU 2L ⊗ U (1)Y Standard Model [34],[35].
Hence taking the decay π 0 → γγ as not making any statement about the
number of colours, the other two experimental evidences above nevertheless
provide unequivocal support to the number of colours being three.
Next arises an important question: Given three colours, why the gauge
group SU (3)c ? Utilising the three colours example, in principle we may have
chosen a gauge group SO(3)c . In SO(3)c the representation of quarks would
be real. So in that case if we have a meson made up of (q~q) then there would
be a corresponding bound state of two quarks (qq). This is true as in this case
the representations are real. But no such 2 − q bound state exists in the quark
model. However, in the SU (3)c case, as the fundamental representation is
complex, the q and q̄ states are distinct from each other. Thus a q q̄ bound state
does not demand a ‘qq’ bound state. If we still gauge the SO(3)c theory, then it
turns out that for up to 6 flavours (u, d, s, c, b, t) and three colours, this theory
does not possess the property of asymptotic freedom. In contrast the SU (3)c
gauge theory (as we discuss below) does indeed exhibit the experimentally
confirmed property of asymptotic freedom for quarks. Others like SU(2) are
ruled out as its three-valued representation (i.e., its adjoint representation ) is
bosonic in nature. The group SU (2) ⊗ U (1) too cannot produce a consistent
antisymmetric wave function. Thus SU (3)c is the only reasonable choice. And
it works!

6.2 QCD Langrangian and Asymptotic Freedom


For the group SU (3)c assume that its fundamental representation [45],
Quantum Chromodynamics 207

 
f1
ψ =  f2  (6.6)
f3
transforms as

ψ 0 (x) = U ψ(x) (6.7)


where

U = exp(iθ~ · T~ ) (6.8)

θ~ = θ~∗ , T~ = T~ †

θ~ · T~ = θ1 T 1 + θ2 T 2 + · · · + θ8 T 8
With the generators of SU (3)c as T~ = ~λ/2 satisfying the Lie algebra

[T j , T k ] = if jkl T l (6.9)
 δjk
T r T jT k =
2
and f jkl are the structure constants of the group SU (3)c .
As a result of the global gauge invariance in Equation 6.7, Noether’s the-
orem demands a conserved current to exist.
λa
J µ,a (x) = ψ̄ (x) γ µ ψ(x) (6.10)
2
such that ∂µ J µ,a = 0.
Now let us go from global transformation to local by making θ → θ(x), i.e.,
let us introduce a dependence upon location x. The transformation Equation
6.7 likewise depends upon x. Thus,

∂µ ψ → ∂µ ψ 0 = ∂µ (U ψ) = U (∂µ ψ) + (∂µ U )ψ 6= U (∂µ ψ)


Hence ψ̄∂µ ψ is no longer an invariant under local gauge transformation
while ψ̄ψ is an invariant. Local gauge invariance is ensured by replacing ∂µ
with Dµ

Dµ = ∂µ − igGµ (x)


= ∂µ − ig ~µ
·G (6.11)
2
Note
208 Group Theory in Particle, Nuclear, and Hadron Physics


→ 8
λ −→ X λa
Gµ (x) = · Gµ = Gµ a (x) (6.12)
2 a=1
2
Thus the 3 × 3 matrix is given as:
G8
 
G3µ + √µ3 G1µ − iG2µ G4µ − iG5µ
G8
 
Gµ (x) = 
 G1 + iG2 − G3 + √µ G6µ − iG7µ 
 (6.13)
µ µ µ 3
G4µ + iG5µ G6µ + iG7
− √23 G8µ
Consequently Gµ transforms as the adjoint representation of the group
SU (3)c .
Thus we take the Lagrangian

L0 = iψ̄γ µ ψ → iψ̄γ µ Dµ ψ (6.14)
∂xµ
where the transformation of Dµ goes as

Dµ → Dµ 0 = U Dµ U −1 (6.15)
Hence

Lint = g ψ̄γ µ Gµ (x)ψ (6.16)


where Gµ (x) is given in Equation 6.13. Then qi qj Gkµ coupling is,

λk
gTji k = g( ) (6.17)
2 ji
Now for the Lagrangian involving the SU (3)c gauge fields alone we define
a second rank tensor as follows:
i
Gµν = [Dµ , Dν ] (6.18)
g
= T~ · G
~ µν (6.19)
where


Gµν = ~ µν
·G (6.20)
2
With
i
Gµν = ∂µ Giν − ∂ν Giµ + gf ikm Gkµ Gm
ν , (i = 1, 2, . . . 8) (6.21)
Then the gauge Lagrangian is
1
Lgauge = − T r(Gµν Gµν )
2
Quantum Chromodynamics 209

8
1 X a µνa
=− G G (6.22)
4 a=1 µν
In contradistinction to the Abelian gauge theories, in case of the non-
Abelian gauge theories – on account of the form of Gµν , the gauge field tensor
– there exist 3- and 4-gauge boson (i.e., gluon) interaction terms in Lgauge as:
0 0
gfabc (∂µ Gaν )Gµ,b Gν,c ; g 2 fabc fab0c0 Gbµ Gcν Aµ,b Aν,c (6.23)
These gluon self-interactions are responsible for most of the difference be-
tween QED (which is a gauged Abelian theory) and QCD here. In QED the
photon does not carry the conserved (electric) charge and thus exhibits no self-
interaction. Consequently it is electrically neutral. This leads to fundamental
differences in how the coupling constants “run” in QED and QCD.
The coupling constants “run” as per the concepts from renormalization
group studies [46]. In Quantum Electrodynamics (QED), built on the Abelian
group U (1)em , Coulomb’s law is corrected by means of vacuum polarization
diagrams. In quantum field theory, the summation of higher order corrections
while retaining only the leading logarithms yields first term in a power series
expansion of α = e2 /~c as:

α(q02 )
α Q2 =

  (6.24)
α(q02 ) q2
1− 3π ln q02

By selecting some arbitrary momentum transfer q02 (denominated the


renormalization point, which may be taken to be q0 = me , the electron mass),
we predict the values of the coupling constant, or its evolution at arbitrary
Q2 .
Thus at shorter distances (or higher values of Q2 ) the effective charge
becomes larger. This is easily understandable in QED. Given a test charge, it
will polarize the vacuum around it and also attract the opposite charge of the
polarized vacuum. So at large distances, the effective charge of the test particle
as per Gauss’s law will be smaller in magnitude. One has to go to distances
sufficiently close to the test charge to feel its “bare” charge. Thus the vacuum
of QED is polarizable. Now QCD, which is built on the foundational structure
developed for QED, albeit for a non-Abelian SU (3)c group, displays certain
similarities to this, in addition to various complexities introduced on account
of its non-Abelian character. One similarity is that Coulomb’s Law in the One
Gluon Exchange is modified by the vacuum polarization diagrams identical
to the QED case. But now in addition the gluons self-couple. This provides
contributions to coupling constants which behave strongly opposite in nature
to the QED-like contributions.
Finally the running coupling constant in QCD is given as
210 Group Theory in Particle, Nuclear, and Hadron Physics

αs (q02 )
αs Q2 =

  2 (6.25)
1+ 11
3 Nc − 32 Nf ln Q
q2
0

Note the difference in signs in Equation 6.25 in QCD with respect to


Equation 6.24 for QED. Nc denotes the number of colours (take Nc = 3), as
long as Nf - the number of flavours does not exceed 16. Experimentally Nf is
6. αs (Q2 ) now for Q2  q02 , goes to zero, meaning that quark-gluon coupling
becomes very small (anti-screening) at large Q2 (i.e., at short distances )
and thus one may take these quarks to be non-interacting in this limit. This
phenomenon of “free” quarks at high momentum transfers is denominated
asymptotic freedom and is a unique stamp of the non-Abelian nature of
SU (3)c .

————————————————–
Unsolved Problem 6.1: SU (3)c gauge group leads to the phenomenon of
asymptotic freedom and this we attribute to the non-Abelian nature. However
just the non-Abelian character of SU 3c in itself is not sufficient to ensure
asymptotic freedom of quarks. As an example, demonstrate that if one gauges
the group SO(3) (a non-Abelian group) then the corresponding system is not
asymptotically free for 6-flavours of quarks.
————————————————–

For SU (3)c QCD on the other hand at scales of Q2 ∼ q02 , for small Q2 (or
large distances) αs (Q2 ) becomes very large. Thus quarks would be strongly
bound and hence confined within hadrons. However, this argument should
be taken as only hinting at confinement of quarks. For this logic in itself
cannot be taken as definitive proof of confinement of quarks inside hadrons,
as the derivation of αs (Q2 ) requires a perturbative series. Doing so cannot be
justified for situations where coupling becomes strong.
The renormalization group concept utilized in deriving α(Q2 ) and αs (Q2 )
depends upon a renormalization point q02 . For a renormalizable theory (as
U (1)em and SU (3)c are) the information about length scale resides within
the renormalized parameters and they vary with a change of scale. Thus a
renormalizable theory describes a phenomena at a particular length scale, in
terms of parameters (e.g., mass and charge in U (1)em ) that can be measured
at that scale only.
Now this book is about group theory and the “renormalization group” ap-
pears to indicate an underlying group structure in what has been performed
so far. However the ideas required here arise from consistent studies in quan-
tum field theory where a differential “flow equation” is required [46] to obtain
renormalized coupling constraints in QED and QCD. This however does not
require any deep-rooted group theoretical concept other than the additive
group of transformation q0 → q0 + δq0 . Thus the word “group” itself in renor-
malization group studies is a misnomer in quantum field theory [47].
Quantum Chromodynamics 211

So is QCD defined by the parameter αs (Q2 )? It cannot be taken as such,


since αs (Q2 ) is a function of the scale q02 at which it is measured. Hence there
is no dimensionless number which we may take as a measure of the strength
of QCD. Let us look at scales at which αs (Q2 ) ∼ 1, i.e., so called ΛQCD .
Here ΛQCD is merely the scale against which other physical quantities may
be measured. Consequently QCD is a zero-parameter theory.

6.3 Colour Singlet States and Confinement


As no free quarks have been identified we have for the colour space:

3 ⊗ 3̄ = 1 ⊕ 8
3 ⊗ 3 ⊗ 3 = 1 ⊕ 8 ⊕ 8 ⊕ 10 (6.26)
Hence it is proposed that quarks are never seen in isolation as these are
bound in the above colour singlet states of mesons and baryons. This is the
so-called colour singlet hypothesis of colour confinement.
So what is the reason for this predilection for colour singlet states over
other colour states in QCD? In order to understand this issue, let us confine
ourselves to the simplest case of one gluon exchange between quarks. Now
we saw that, although gluons in QCD are strongly self-interacting systems,
these ultimately turn out to be mediators of interquark forces. In fact, in the
quark model (as we see in the next chapter) one gluon exchange interaction
provides a very successful picture of a potential acting between quarks. Thus
we accept, on phenomenological grounds, that one gluon exchange interaction
is a valid potential to work with. The analysis below, which provides the
relative strength of different representations, is termed “Maximally Attractive
Channel” ([48] p.120).
Let the three colours in SU (3)c be denominated R, B, G. The octet of the
gluons is given as:

RB̄, RḠ, B R̄, B Ḡ, GR̄, GB̄,

RR − B B̄ RR̄ + BB − 2GḠ
√ , √ (6.27)
2 3
The last two are colour-preserving gluons which are orthogonal to the
excluded colour singlet combination
1
√ (RR̄ + B B̄ + GḠ) (6.28)
3
The elementary interaction between quarks is say
212 Group Theory in Particle, Nuclear, and Hadron Physics

(blue − quark) + (RB̄ − gluon) → (red − quark) (6.29)


Let us consider the interaction between two G-quarks mediated 
by (RR̄
+
√ 2
B B̄ − 2GḠ)/ 3. Then the interaction strength is proportional to − √6 ×
 
− √26 = + 23 . Hence, as is the case of electromagnetism for like-charged
electrons, identical-colour charges in QCD repel.
Next interaction between G-quark  and
 B-quark by direct interaction
√ 
through (RR̄ + B B̄ − 2GḠ)/ 3 gives √6 × − √6 = − 13 and an exchange
1 2

term through GB̄ gluons is 1 × 1 = 1. Total strength of this interaction is


− 31 + 1 = 23 .
Now two quarks combine to give a symmetric combination, say S =
GB + BG and an antisymmetric combination A = GB − BG. The direct
term gives the interaction energy as (S + A) − 31 and the exchange term
contributes (S − A) (1). Thus interaction strengths of the symmetric and the
antisymmetric terms are, respectively,
2 4
es = and es = − (6.30)
3 3
These results are independent of the labels R, B, and G.
So for three quark-colour combinations, states are obtained by counting
the number of symmetric and antisymmetric pair exchanges involved in a
particular representation as follows:
(a) Totally antisymmetric colour state

with interaction strength

e1 = 3edirect + eexchange(A) = 3eA = −4 (6.31)


(b) Totally symmetric colour state

with interaction strength

e10 = 3edirect + eexchange(S) = 3eS = +2 (6.32)


(c) The colour octet state,
Quantum Chromodynamics 213

having interaction strength as

 
1
e8 = 3edirect + eexchange(S) + eexchange(A) = 3 − + 1 − 1 = −1 (6.33)
3

While (b), the symmetric state, is repulsive, the singlet state (a) is the most
attractive. Hence as per maximally attractive channel analysis the three-quark
colour singlet is the most attractive and hence the preferred configuration for
the ground state.
Thus the maximally attractive channel analysis, being so successful in
providing a highly plausible explanation for the stability of the colour singlet
state for 3-quarks, may be generalized in a straightforward manner to that of
nq -quarks which exist in a representation with nS number of symmetric pairs
and nA number of antisymmetric pairs having an interaction strength given
as [49],[48]:
1  nq 
e = (nS − nA ) − (6.34)
3 3
where
n  nq !
q
= (6.35)
2 (nq − 2)!2!
Example: Take the case of 6-quarks in the following representations

 
1 6
; e=3−6− = 3 − 6 − 5 = −8
3 2

; e = 4 − 4 − 5 = −5

; e = 6 − 3 − 5 = −2

; e = 6 − 3 − 5 = −2

; e=7−2−5=0
214 Group Theory in Particle, Nuclear, and Hadron Physics

The first one is preferred as the maximally attractive channel. However, it


is exactly equal to a pair of 3-q each in an antisymmetric channel,

= + = −4 − 4 = −8
And thus there exists no extra attraction for the case of 6 − q vis-a-vis
3q + 3q antisymmetric channels and thus 6-q is not necessarily more bound
than the 2-colour singlet 3-q states. Hence the 6-q colour singlet state would
split instantaneously into two free and independent 3-q colour singlet states.
This speaks against the existence of dibaryons like H-dibaryons (which have,
in spite of intense searches, not been found in the laboratory as of now).

————————————————–
Problem 6.1: Calculate the 9-q, 12-q, and 15-q colour singlet states as
per the maximally attractive channel formalism. What can one conclude for
these states?
————————————————–

Let us look at the mesonic states of the quark-antiquark pair. Consider√q q̄


in the colour octet ḠR state. The direct interaction term (RR̄+B B̄−2GḠ)/ 6
gives the strength as
 √   √  1
e8 = 1/ 6 + (−) −2/ 6 =
3
This is repulsive and hence the octet state in 3 × 3̄ would not be bound.
Now for the colourless state, take ḠG state. Here there are these possible
interactions:

(a)
√G √R̄+ B B̄2 − 2GḠ)/ 6 if it preserves colour is e1 =
 Ḡ with (R
−2/ 6 (−1) −2/ 6 = − 3

(b) ḠR gluon would give GḠ → RR̄ transition as e(2) = 1 (−1) = −1 and
ḠB gluon to GḠ → B B̄ case as e3 = 1 (−1) · 1 = −1

Now the meson exists in a colour singlet state as (R̄R + B̄B + ḠG)/ 3.
We notice that the ḠG state goes to ḠG,√R̄R, and B̄B through terms like
e1 , e2 and e√
3 . Hence ḠG weighted by 1/ 3 projects onto the state (R̄R +
B̄B + ḠG)/ 3 and as there are three such colour terms, this yields the total
interaction strengths for q q̄ SU (3)c . Thus,
8
esinglet = e(1) + e(2) + e(3) = −
3
This is an attractive colour singlet channel for mesons to exist in. This is
Quantum Chromodynamics 215

essentially in accord with what we expected from Equation 6.26, and thus en-
hances our confidence in the correctness of the logic inherent in the maximally
attractive channel analysis.

6.4 Pentaquarks
Baryonic (QQQ) and mesonic (qq) configurations are known to exist in
colour singlet states. Thus exotic [(qqq) (qqq)] quark states may exist. Indeed
they play a significant role in the structure of A=2 baryons (as we discuss
in Chapter 12) as do [(qqq) (qqq) (qqq)] and [(qqq) (qqq) (qqq) (qqq)] states in
A=3 and A=4 nuclei, respectively. But these exotics only play a role deep
inside the A=2,3,4 nuclei, as shall be elucidated in Chapter 12.
QCD permits the existence of exotic configurations such as [(qq) (qq)] and
[(qqq) (qq)]. The latter is the so-called pentaquark state. This pentaquark,
comprised of four quarks plus one antiquark, has been the focus of much
theoretical and experimental research.
It has been claimed that, as the proton is composed of three quarks (uud)
and as there exists a veritable sea of (qq) pairs that is evident from neutrino
and antineutrino scattering experiments, one may hence assume the proton
to have (qqq) (qq) states, namely pentaquark configurations, built into it.
But the alert reader should see the falseness of this logic. For the above
quarks in (qqq) (q q̄) are all current quarks, and as such what the experiments
in actuality demonstrate is that there are three-valence current quarks and a
sea of qq as (q q̄ q q¯ q q̄ . . . ). The sea of q q̄’s cannot be converted into a single
pair of (q q̄). And hence there is no such configuration as (qqq) (q q̄) in the
proton. In fact, as per our discussion of the current quark and the constituent
quark, the three valence current quarks may be treated as three constituent
valence quarks with each current quark surrounded by the sea of (q q̄) plus
gluons.
If the pentaquark is composed of u−d flavours, such as uuddd, ¯ only, then it
would be difficult to separate it from (uud)+(dd) ¯ configurations. However with
strangeness S = +1 as in uudds, it may be identifiable by using conservation
of strangeness. The standard baryons like Λ (uds) and Σ+ (uus) exhibit the
S=–1 configuration which would be distinguishable from the exotic S=+1
state of the pentaquark (uudds̄).
Consequently the pentaquark is constituted of four quarks and an anti-
quark. So what does the maximally attractive channel method (which has
shown to be successful in determining the stability of colour singlet channels
for qqq, (q q̄) and multiquark systems) have to say about the pentaquark? Us-
ing Equation 6.34 and assuming that in the 4-q system, the 3-q are already
in the colour singlet state:
216 Group Theory in Particle, Nuclear, and Hadron Physics

 
1 4
; e = (1 − 3) − = −4 (6.36)
3 2
which is the same as the energy of a quark and a proton or neutron separated
as

+ ; e = −4 + 0 = −4 (6.37)
So here, there is no indication of any extra binding. For a similar situation in
the 6-q case we determine that the totally antisymmetric state is not any more
bound than two separated 3-q nucleons and concluded that, if found, the (6-q)
state would instantaneously split up into two (3-q) colour single states. Hence
on the basis of similar analysis we conclude that the 4q would instantaneously
split up locally into a (3-q) colour singlet and a 3-plet colour quark.
Let us study all the possible representations for the 4-q states in SU (3)c
space. So given

⊗ ⊗ = ⊕ ⊕ ⊕

( ⊗ ⊗ ) ⊗ = ( ⊕ ⊕ ⊕ ) ⊗

= ⊕( ⊕ ⊕ )twice + ( ⊕ )
Let us use maximally attractive channel analysis for all the Young Diagram
representations:

 
1 4
; e=1−3− = −4
3 2

; e=3−1−2=0
Quantum Chromodynamics 217
TABLE 6.1: Hidden colour components in the
multiquark systems
Multiquark configurations Percentage of hidden colour
qq q̄ q̄ (qq in [3̄]) 33
qq q̄ q̄ (qq in [6]) 66
Pentaquark qqqq q̄ 66
A = 2 ; 6-q 80
A = 3 ; 9-q 97.6
A = 4 ; 12-q 99.8

; e = 2 − 2 − 2 = −2

; e=6−0−2=4
Thus the first one is the most attractive representation. So as we saw above
for this representation, the maximally attractive channel analysis does not
favour the formalism of 4 − q tagging on to a q̄ to form a pentaquark.
We have hidden colour components of 6-q, 9-q, and 12-q states as discussed
in Chapter 12. There we noticed the repulsive character of hidden colour
components for nuclear A = 2 , 3 and 4 systems, respectively. In Table 6.1
we list the hidden colour components of these plus some other multiquark
systems and pentaquarks [50].
Suppression of large hidden colour components in hadrons would thus dis-
courage formation of pentaquarks and other exotic hadrons.
Hence we notice that both the hidden colour concept and the maximally
attractive channel analysis speak against the formation of pentaquarks. This
may explain why it has been so hard to see pentaquarks in the laboratory.
But is there a catch to the above analysis? In the maximally attractive
channel analysis, we look at the SU (3)c colour representations only, and thus
the method is independent of the flavour degree of freedom. Thus here the
baryons may be made up of (uud)-quarks or (uuc)-quarks (with c-quark in
SU(4)) and the conclusions drawn from the maximally attractive analysis
would hold for both. However the hidden colour analysis has been performed
for the group structures:
SU (18)F SC ⊃ SU (6)F S ⊗ SU (3)c ⊃ SU (3)F ⊗ SU (2)S ⊗ SU (3)c
So for up to 3-flavour (u,d,s) of quarks, the hidden colour analysis dis-
favours formation of exotics, including the pentaquark. But note that it as-
sumes that SU (18)F SC is a good group with the three small quarks providing
a valid fundamental representation of this large group. For the three light
quarks the above appears to be a reasonably good assumption.
However take four flavours (u, d, s, c) of quarks. Now, is it that we can
generalize to a larger group decomposition?
SU (24)F SC ⊃ SU (8)F S ⊗ SU (3)c ⊃ SU (4)F ⊗ SU (2)S ⊗ SU (3)c .
218 Group Theory in Particle, Nuclear, and Hadron Physics

A pentaquark (uudds̄) with S = +1 called θ+ was claimed to have been


discovered in 2002 by a group in Japan. This was ∼1.5 times heavier than the
proton. Within a year, more than 10 other groups confirmed the finding of this
putative θ+ pentaquark by reanalyzing their data. But a few others did not
observe the state. However in 2005 Jefferson Laboratory, repeating the same
experiment as the Japanese group, conclusively ruled out the very existence
of the θ+ pentaquark. Hence the θ+ pentaquark was a mirage. The Particle
Data Group (PDG) in 2008 described the above pentaquark pseudo-discovery
as, “a curious episode in the history of science”.
However, the lure of the pentaquark for experimentalists has continued.
Only recently the LHCb experiment at CERN [51] has made a strong claim of
the discovery of a pentaquark – actually of two states at 4.38 and 4.45 GeV
− +
with preferred spins of 32 and 25 , respectively. This pentaquark is constituted
of {(uud) cc̄}, i.e., no s-quark, but a charm-anticharm pair with (uud).
A statistical scale of 5-sigma is normally considered to be sufficient re-
quirement for a claim of new particle discovery to be taken seriously. In this
case it was 9-sigma! Hence we take this discovery of pentaquark seriously. But
the issue is even more complicated in that, instead of a single pentaquark, we
have here a pair to contend with.
Note that the demand for the multiquark states (like pentaquark
(qqq) (q q̄)) to be colour singlets, together with the Pauli exclusion principle,
leads to the possibility of assigning these states to be a completely antisym-
metric representation of the group:

SU (18)F SC ⊃ SU (6)F S ⊗ SU (3)C ⊃ SU (3)F ⊗ SU (2)C ⊗ SU (3)C (6.38)


This is true of three flavours (u, d, s) of SU (3)F .
The Pauli principle when applied to these multiquark states demands
proper embedding of the SU (6)F S ⊗ SU (3)C group in SU (18)F SC . Then only
can one obtain unique states with proper hidden colour components. Hence
in 4-q when the fourth quark is brought on to the colour singlet (3-q) system,
the fourth quark should observe the Pauli Exclusion Principle with respect to
other quarks. If we can somehow ensure that this fourth quark is not subject
to the Pauli Exclusion Principle with respect to the other three, then we may
be able to bypass the hidden colour constraints. But how can this apparent
contortion of the basic laws of nature be?
Let us seek help from hypernuclear physics. A hypernucleus is created
when a Λ hyperon is captured by a nucleus consisting of Z number of protons
and N number of neutrons. Now the protons and neutrons in a nucleus are
subject to the Generalized Pauli Exclusion Principle wherein exchange of two
nucleons yields an antisymmetric function. Hence, for example, if one adds a
nucleon to a closed shell 4 He nucleus with (p↑ p↓ n↑ p↓ ) filling the lowest orbital,
then this nucleon, due to the Pauli Exclusion Principle, must go to the next
higher orbital. However the Λ-hyperon is not subject to the Generalized Pauli
Exclusion Principle with respect to the other nucleons, and thus sinks to the
Quantum Chromodynamics 219

bottom of the nuclear shell. In other words, it goes right to the very centre
of the nucleus. This is an empirically well-confirmed property of the light and
the medium mass hypernuclei.
Note that in SU (6)SF ⊗SU (3)c the colour space is antisymmetric for three
quarks, and the SU (6)SF ⊃ SU (3)c ⊗ SU (2)S component is symmetric, thus
yielding antisymmetry in total. This is good enough for three quarks but when
extended to multiquark configurations to ensure proper antisymmetry, extra
caution is required. This is demonstrated in detail in the hidden colour section
and has been amply discussed by Matveev and Sorba [52]. It is necessary to
resort to the SU (18)SF C group, and as in Equation 6.38 only by means of the
proper embedding of subgroups is it possible to correctly obtain symmetrized
functions for multiquarks.
Next take q ∼ (u, d, s, c) as the fundamental representation of SU (4)F . But
now as mc ∼ 1.5Gev  ms , md , mu so this symmetry is badly broken. Then
how are we still able to build (qqq) in SU (4)F as a properly symmetrized state
in SU (8)F S to go with the SU (3)c antisymmetric state? One may conjecture
that this may be due to the strong antisymmetry arising from SU (3)c . This
then ensures the proper symmetric state in SU (8)F S for (qqq) charm baryons.
But this is not good enough for the multiquarks with a c-quark. Inclusion of
c-quarks requires the larger group structure

SU (24)F SC ⊃ SU (8)F S ⊗ SU (3)C ⊃ SU (4)F ⊗ SU (2)S ⊗ SU (3)C (6.39)

where the 24-dimensional fundamental representation plays a basic role in


ensuring proper symmetry of the multiquark states. Now, SU (4)F symmetry
being pretty badly broken as it is, there is no way that SU (24)F SC may be
visualized to be a relevant symmetry at all. Note that the colour merging
with 4 flavours and 2 spins in the 24-dimensional fundamental representation
of SU (24)F SC loses its power, which it had for the 3-quark baryons (as shown
above) and thus this larger group has no relevance for charm-quarks as part
of a multiquark state.
So the only physically acceptable manner that the c-quark can be a mem-
ber of the multiquark state is by not being identical to the (u, d, s) quarks of
SU (3)F . Therefore, no Pauli Exclusion Principle applies for the c-quark with
respect to exchange with the (u, d, s)-quarks in a multiquark system.
Thus the situation is very similar to that of the Λ- hyperon with respect to
the (p,n) based nuclei. Now a single c-quark entering a bound (uud) system
not partaking in the Pauli Exclusion Principle will go to the lowest orbital
and reside at the centre of this 4-quark system. We may visualize this as a
solitary c-quark surrounded by a “cloud” of colour singlet (uud) quarks. This
may be treated as a quasi-c-quark. Now it seeks a c̄ to form a bound state as
(uudc)c̄ in a colour singlet state of 3X 3̄ = 1 + 8.
The J/ψ state of cc̄ is J p = 1− and has a mass of 3.097 GeV (c-quark has
a mass mc ∼ 1.5GeV ). What state does the (uud) colour singlet configuration
exist in to provide the cloud for the c-quasi-quark state? As per the Particle
220 Group Theory in Particle, Nuclear, and Hadron Physics

Data
 Group (PDG) (2012)  [53] for (uud) we havethe following  possibilities:
+ + −
p J p = 21 , N (1440) J p = 12 and N (1520) J p = 32 .
Not expecting the (uud) to be in the ground state, we exclude protons and
thus the other two may attach to a c-quark with spins:
+ − − −
(1) 12 + 12 = (1 + 0)+ which with c̄ would give (1 + 0)+ 12 = 32 + 21 ,

h i−
− − − + + + +
(2) 32 + 12 = (2 + 1) with c (2 + 1) + 12 = 25 + 32 + 32 + 12 . (Note
extra negative parities as cc̄ is attaching). We use the maximal stretched state
− +
to obtain 32 and 25 spins. This seems to be similar to the nuclear spin-orbital
− →
→ − →
− → − 3
force for which, e.g., 1 + 12 = 32 + 12 makes p 2 lower than the p 21 states.
h i 

So 4.38 GeV 23 state is made up of (uud)N (1440) + c +c
quasi
h i 
5 +
and 4.450 GeV 2 states is constituted as (uud)N (1520) + c +c .
quasi

Hence this picture not only provides a convenient visualization as to what


this uudcc pentaquark is, it is also able to explain why there should be two
associated states.
Given the fact that the Υ meson at 9.460 GeV with J p = 1− should be
accompanied by a similar (uudbb) pentaquark just above its mass of 9.46 GeV,
a clear prediction of this model is the existence of a pair of states.

6.5 Large Nc Colour QCD


As we saw, QCD is a zero-parameter theory. Hence we cannot perform per-
turbative calculations at large distances. Hence one may take Nc , the number
of colours, as a hidden “weak-coupling” parameter. Though Nc = 3 is an ex-
perimentally verified number, however, one may assume Nc as an arbitrary
parameter and develop the theory accordingly. After completion of necessary
analysis, one can then return to Nc = 3 to match experimental results. We
need to study Nc as the free (large) parameter in QCD based on the group
SU (N c ), as it turns out that 1/Nc expression can be taken to provide us with
a perturbatible series. This study of QCD in large Nc limit has proven to
be an eminently fruitful theoretical enterprise. While the number of quarks,
as the fundamental representation of SU (Nc ), scales as ∼ N c , the number of
gluons, belonging to the adjoint representation, scales as ∼ (Nc2 − 1). So for
large Nc , gluons will dominate over quarks. It also turns out that the field the-
ory of SU (Nc ) for large Nc reduces to a theory of weekly interacting mesons.
One thus connects this theory to the older topological Skyrme model where
baryons arise as topological structures in a Lagrangian composed of scalar
Quantum Chromodynamics 221

mesons only. We shall study this in more detail in Chapter 14. Here we wish
to study how large Nc QCD may provide useful properties of baryons.
One useful property of baryons composed of Nc -quarks in SU (Nc ) is that
their mean square radius is independent of Nc [54]. Also their mass goes as
mβ ∼ Nc , i.e., the number of colours. Only quarks in such a system can bind
to baryons. So if the mass of the baryon turns out to scale as mβ ∼ Nc 2 , then
the system will be unbounded. These properties of quarks in SU (Nc ) QCD
actually match the structure of baryons in the Skyrme model (Chapter 14).
So in the SU (Nc ) model baryons are composed of Nc number of quarks.
However if Nc were an even number, say Nc = 4, then the composite baryon,
being built out of an even number of spin − 21 entities, shall have a total spin
of 0 or an integer. But we want our baryon to be a fermion and hence we
demand that Nc be an odd number like Nc = 1, 3, 5, 7 . . . . To ensure this take
[54],

Nc = 2k + 1, k = 0, 1, 2, 3 . . . . (6.40)
Now assume that the proton is built up of (k + 1) number of u-quarks and
k number of d-quarks. And vice-versa for the neutron. Now Witten et al. [54]
took quark charges to be the same for any arbitrary Nc as

Qu = 2/3 and Qd = −1/3 (6.41)


So their charges were static charges, independent of any colour, and so remain
the same for any arbitrary Nc colour. Thus in their model the proton and
neutron charges are
   
2 1 k+2 Nc + 3
Qp = (k + 1) + k − = = (6.42)
3 3 3 6
 
2 1 k−1 Nc − 3
Qn = k + (k + 1) (− ) = = (6.43)
3 3 3 6
Now for Nc = 3 these gave proper charges Qp = +1 and Qn = 0. However
for arbitrary Nc , the results are not even integral. For Nc = 5 , Qp = 4/3, Qn =
+1/3 these charges actually blow up as Nc → ∞. This is an unsatisfactory
behaviour of static charges in the models of Witten et al. for QCD with large
Nc values. In fact it is catastrophic for the following reason.
For protons and neutrons as bound states of Nc quarks, there exists the
basic Coulomb self-energy term, which will contribute to the total mass of
this bound system. The Coulomb self-energy of the above proton (built up of
static charges of Witten et al. [54]) would be:
2 2
C[(k + 2)/3] C[(Nc + 3)/6]
Ec = = (6.44)
R R
where R is the size of the charge distribution and C is the factor depending
upon the exact shape of the charge configuration. Note that this energy blows
222 Group Theory in Particle, Nuclear, and Hadron Physics

up as Nc squared. Now we expected mB ∼ N c . But in addition we have this


Coulomb self-energy term dominating as Nc 2 and blowing up. This is dis-
astrous for the model of Witten et al. [54] and hence this faulty framework
should be discarded.
Next let us quote our result of colour-dependent electric charges as ob-
tained in the SU (Nc ) ⊗ SU (2)2 ⊗ U 1Y Standard Model as [34],
 
1 1
Qu = 1+
2 Nc
 
1 1
Qd = −1 + (6.45)
2 Nc
New proton and neutron charges in the SU (NC ) model are
   
1 1 1 1
Qp = (k + 1) 1+ +k −1 +
2 Nc 2 Nc
(2k + 2) (−2k)
= (k + 1) +k =1 (6.46)
2(2k + 1) 2(2l + 1)
   
1 1 1 1
Qn = k 1+ + (k + 1) −1 + =0 (6.47)
2 Nc 2 Nc
And hence Qp = 1 and Qn = 0 for any arbitrary Nc , i.e., it is independent
of Nc . Also the Coulomb self-energy term of the proton remains finite. Thus
the colour-dependent electric charge of the SU (Nc ) ⊗ SU (2)2 ⊗ U (1)Y model
successfully takes care of the problem plaguing the static quark charge frame-
work of Witten et al. [54]. This demonstrates that these colour-dependent
charges in Equation 6.45 are the correct charges to be taken in the studies of
quarks in SU (Nc ) QCD [34].
In SU (Nc ) ⊗ SU (2)L ⊗ U (1)Y spontaneously broken by the Englert-Brout
mechanism to SU (Nc ) ⊗ U (1)em , the final unbroken gauge symmetries are
QCD and QED. Now clearly QED is independent of QCD, and the photon
does not know of colour and does not interact with gluons. But surprisingly
the electric charge of quark (Equation 6.45) has contributions from colour
(Nc )! Thus the electric charge of the quark obtains contributions from strong
interaction as well. This manifestation of colour in the structure of the electric
charge is a sign of a ‘new’ unification in the Standard Model built on the group
structure SU (Nc ) ⊗ SU (2)L ⊗ U (1)Y . In fact the electro-weak unification is
possible only because there is a pre-existing colour group interwoven in the
fabric of the larger Standard Model. What one is saying is that we cannot
treat the electro-weak group SU (2)L ⊗ U (1)Y in isolation and hope to obtain
a complete description of the electromagnetic and weak interaction effects.
We have to include the colour group SU (3)c as well. Thus the unified and
complete model is the whole group structure SU (Nc )⊗SU (2)L ⊗U (1)Y . Hence
the word ’Standard Model’ is not merely a convenient expression to describe
a mixed structure of two separate realities of the strong interaction joined
Quantum Chromodynamics 223

together from the outside with the electro-weak interaction. This appears
to be suggested by the group structure SU (Nc ) ⊗ SU (2)L ⊗ U (1)Y being a
product of different groups. But these, in particular the anomalies, bring in
an intrinsic unity between the quarks and the leptons which mix the colour
with the electric charge! So all, the strong, the electromagnetic and the weak
forces are already unified in the Standard Model. One does not have to resort
to the fanciful Grand Unified Theories in order to obtain some hypothesized
putative unification. It is already present in the Standard Model. Empirically
the SU(5) GUT has been ruled out anyway [55].
Now let us calculate the magnetic moment of these baryons in SU (Nc )
QCD. Now we know that in proton we have (k + 1) and k quarks of u- and d-
~u = (k+1) and S
flavours, respectively. Let us assume that their spin is S ~d = k
2 2
and then S ~p = S
~u + S~d = . 1̄
2
Take the magnetic moment operator as
1 ~u + µd S
~d
µ
~ = µu S (6.48)
2
where µq = Qq · e~/(2mq c) (note µq = 12 gq in Dirac theory). The quark
charges are the correct colour-dependent charges in Equation 6.45.
Next now

~ = Sz {2µu [Su (Su + 1) + S (S + 1) − Sd + 1] + (u ↔ d)} /2S (S + 1)


µ
(6.49)
and so
1 k+1 k+1 1 1 k k
µ(p) = [2µu [( )( + 1) + ( + 1) − ( + 1)]
2 2 2 2 2 2 2
k k 1 1 k+1 k+1 1 1
+2 µd [ ( + 1) + ( + 1) − ( )( + 1)]]/[2 + ( + 1)]
2 2 2 2 2 2 2 2
1
= [(k + 3) µu − k µd ] (6.50)
3
And then u ⇔ d gives
1
µ(n) = [(k + 3) µd − k µu ] (6.51)
3
Next using Equation 6.48 as ∆+ is also built of (k + 1)-u quarks and k-d
quarks but with S = 3/2 :
3 k+1 k+1 3 3 k k
µ(∆+ ) = 2 · [µu [ ( + 1) + ( + 1) − ( + 1)]
2 2 2 2 2 2 2
k k 3 3 k+1 k+1 3 3
+µd [ ( + 1) + ( + 1) − ( + 1)]]/[2 · ( + 1)]
2 2 2 2 2 2 2 2
1
= [(k + 9)µu − (k − 6) µd ] (6.52)
5
224 Group Theory in Particle, Nuclear, and Hadron Physics

One finds that these magnetic momenta are also very successful for the
SU (Nc ) model [34].

————————————————–
Problem 6.2: Taking the electric charges of the six-flavors (u,d), (c,b)
and (t,b) as belonging to the three generations in the Standard Model as
 
1 1
Qu = Qc = Qt = 1+
2 Nc
 
1 1
Qd = Qs = Qb = −1 + (6.53)
2 Nc
show that the charges of all the baryons in the spin-1/2 octet and the spin-3/2
decuptet evaluate correctly for any Nc in the QCD model.
————————————————–

6.6 Solutions of Problems

Solution 6.1:
(a)

 
1 9
; e=9−9− = −12 = 3(e3q )antisymmetric
3 2

(b)

 
1 12
; e = 3 × 6 − 12 − = −16 = 4(e3q )antisymmetric
3 2

(c)

 
1 15
; e = 30 − 15 − = −20 = 5(e3q )antisymmetric
3 2
Quantum Chromodynamics 225

So none of these are more attractive than the appropriate number of indi-
vidual colour singlet 3-quark entities.
————————————————–
Solution 6.2:

1 1 1 1
Q∆0 = kQu + (k + 1)Qd = k (1 + ) + (k + 1) (−1 + )=0
2 Nc 2 Nc

Q∆++ = (k + 2)Qu + (k − 1)Qd = 2


Q∆− = (k − 1)Qu + (k + 2)Qd = −1
QΣ+ = QΣ+∗ = (k + 1)Qu + (k − 1)Qd + Qs = +1
QΛ = QΣ0 = QΣ0∗ = kQu + kQd + Qs = 0
QΣ− = QΣ−∗ = (k − 1)Qu + (k + 1)Qd + Qs = −1
QΞ0 = kQu + (k − 1)Qd + 2Qs = 0
QΞ− = (k − 1)Qu + kQd + 2Qs = −1
————————————————–
Chapter 7
Quark Model

7.1 Current and Constituent Quarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227


7.2 The Eightfold Way Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
7.3 SU (3)F Flavour Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
7.4 Quark Model Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
7.5 SU (6)SF Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259
7.6 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264

7.1 Current and Constituent Quarks


In the previous chapter we saw how Quantum Chromodynamics (QCD)
successfully explains hadronic structure at high energies. In that framework,
the three quarks in a spin 21 baryon are treated as massless and point-like
particles. These are denominated as current quarks.
In contrast, at low energies theoretical consistency demands that the
quarks be treated as possessing a finite size (i.e., be non-point-like structure)
and a finite mass (about 300 MeV). These are then termed as constituent
quarks. Hence, the proton is comprised of three such constituent quarks; two
u-quarks and one d-quark.
This complete change in the fundamental character of quarks, from con-
stituent quarks at low energies to current quarks at high energies, is believed
to arise from confinement. This characteristic was also discussed in previous
chapters.
Now, current quarks are those which arise in the Lagrangian of the the-
ory of strong interactions – QCD. These are point-like and massless, or of a
very small mass (∼ 4-8 MeV for u- and d-quarks). Experimentally, current
quarks are studied in deep inelastic scattering, and are specified by quantum
numbers such as electric charge, baryon number, weak hypercharge numbers
and isospin. Hence these are the quarks which arise in the Standard Model of
particle physics.
In contrast, constituent quarks possess the quantum numbers which cur-
rent quarks have, but are much heavier ∼ 330 MeV for the u- and d-quarks.
Also these are not point-like and have a finite size of ∼ 0.3 fm inside the
hadrons. These properties manifest themselves at low energies and are thus
used in various quark models.

227
228 Group Theory in Particle, Nuclear, and Hadron Physics

The relationship between these two types of quarks manifests itself in the
laboratory in a bipolar manner: current quarks exist at high energies, while
constituent quarks are found at low energies. Now, as these two extremely
divergent pictures specify the self-same entity, the natural question arises as
to how these contrasting visualizations are related. Conventionally, a hand-
waving argument is applied, treating the constituent quarks as quasi-particles
which may, at low energies, be viewed as current quarks surrounded by an
immense sea of quarks, anti-quarks, and gluons. While easy to visualize, this
depiction is however difficult to implement in a definite mathematical model.
Meanwhile, constituent quarks are phenomenological model entities which
are exactly specified by their masses and dimensions. In the conventional
picture, these two physical properties, mass and size, are the only physically
relevant quantities which distinguish current quarks from constituent quarks.
Is it possible to ascertain further physical properties which can serve to
differentiate current quarks from constituent quarks? We shall define some
new physically measured quantities which will be demonstrated to provide
definite novel physical insights regarding the structure of hadrons in various
quark models.
In deep-inelastic lepton-hadron scattering cross-sections it is observed that,
in contrast to the elastic cross-sections, momentum transfers at high energies
are large and only weakly q 2 dependent (where q 2 is the squared momentum
transfer). This is generally taken as a signal of elastic scattering from point-
like constituents inside the nucleon [45], which can be identified with quarks
and gluons (also denominated partons).
Now let us imagine an “infinite momentum frame” where the target proton
has a very large 3-momentum and let us assume a single photon exchange.
We can neglect the photon mass and so it has 4-momentum P = (ip, p ,0,
0). We visualize these as a parallel stream of quasi-free partons, each with
4-momentum xp, where 0 < x < 1 (see Figure 7.1).
If P is large, we neglect the masses and transverse momentum components
of the partons. Assume that one of the three partons of mass m in the proton is
scattered elastically by absorbing the 4-momentum q of the scattered lepton.
So [42]

(xP + q)2 = −m2 = 0


if
2 2
x P = x2 M 2  q 2

q2 q2
x=− = (7.1)
2P q 2M ν
Here the invariant scalar product P q remains to be evaluated. In the labo-
ratory frame, the energy transfer is ν and for a nucleon at rest, x represents the
fractional momentum of partons in the infinite momentum system. Imagine in
Quark Model 229

e
e
q xP

(1-x) p
proton

FIGURE 7.1: Infinite momentum frame in deep inelastic scattering

the laboratory system, the free point-like particle of mass m,pin an elastic col-
lision receives ν as energy transfer and 3-momentum q 0 = (ν + m)2 − m2 ,
then the 4-momentum transfer is given by the elastic relation,
2
q 2 = q 0 − ν 2 = 2mν (7.2)
Thus

q2 m
x= = (7.3)
2mν M
This we interpret to mean that x is the fractional mass of the nucleon
carried by the free parton which was initially at rest in the laboratory [42].
So for x= 13 the mass of a single quark is m ∼ 310 MeV. Thus here for x= 13
we may attribute this to the current quark with m=0 at x=0 having become a
constituent quark of mass ∼ 310 MeV. Hence consistency demands that only
the single point x = 31 defines the constituent
Rx quark’s mass. R
x
Now let us turn to the case of x= 0 dx. So the integral 0 dx yields x
R1 R1
and hence we determine that 0 dx = x = 1. Thus the full integral 0 goes
beyond the non-relativistic quark model which is characterized by mq ∼ 13 M
(for x; 0 → 13 ). This is the constituent quark mass.
R1
Now let us study the operator < ô >= 0 ôdx. So what we are getting
230 Group Theory in Particle, Nuclear, and Hadron Physics

is a “quasi-quark” which exists beyond the constituent quark of the non-


relativistic quark model.
R 1 This distinctive quasi-quark entity arises from the full integration < ôM>=
0
ôdx, while constituent quarks (of the NRQM) originate from mq = 3 ∼
310 M eV with variable x integrated up to 31 only.
Thus this quasi-quark is a more comprehensively generalized entity mathe-
matically arising from a full x-integration. Hence it should contain constituent
quarks as a special case.
In deep-inelastic electron-nucleon scattering we define the structure func-
tion F2ep (x) for the quark density weighted by the square of quark charges
[42].

F2ep (x) 4 1
dp (x) + d¯p (x) + sp (x) + s̄p (x)

= [up (x) + ūp (x)] + (7.4)
x 9 9
where u(x)dx is the number of u-quarks in proton that carry a fractional
momentum in the range x → x + dx. Similarly for antiquarks and other
quarks as well.
Electron-neutron scattering data is obtained by comparison of electron-
proton and electron-deuteron scattering. Isospin invariance insists that the
u, ū population in a neutron be equal to the d, d¯ population in a proton and
vice versa.

up (x) = dn (x) = u(x); dp (p) = un (x) = d(x); sp (x) = sn (x) = s(x) (7.5)

Therefore

F2en (x) 4 ¯
 1
= d(x) + d(x) + [u(x) + ū(x) + s(r) + s̄(x)] (7.6)
x 9 9
(p+n)
Take a nucleon as N = 2 , that is an equal admixture of neutrons and
protons.

F2eN (x) 5  ¯
 1
= u(x) + ū(x) + d(x) + d(x) + [s(x) + s̄(x)] (7.7)
x 18 9
Next take,

q(x) = qV (x) + qS (x) (7.8)


where V and S stand for valence and sea, respectively.
The difference between the proton and the neutron structure functions is
1
F2ep (x) − F2en (x) =
x[uV (x) − dV (x)] (7.9)
3
So the sea of q q̄ does not contribute here, and only the valence quarks
Quark Model 231

play a basic role. The experimental data for the above expression displays a
dramatic peak at x ∼ 31 [[42], p.231] and thus agrees with it.
For neutrino scattering on nucleons, as there is no electric charge, one
needs two structure functions.
F2νN (x) ¯
= u(x) + d(x) + ū(x) + d(x) (7.10)
x
¯
F3νN (x) = u(x) + d(x) − ū(x) − d(x) (7.11)
Let us assume that the s-quark contribution is small. In that approxima-
tion,
5 eN
F2eν (x) =
F (x) (7.12)
18 2
This agrees well with experimental results and is the best evidence that
nucleons contain fractionally charged quarks.
From the above we see that
Z 1
¯
 
x u(x) + ū(x) + d(x) + d(x) dx
0
Z Z
18
= F2νN (x)dx ≈ F2eN (x)dx (7.13)
5
And this is equal to ∼ 0.50. So the total momentum fraction carried by
quarks and antiquarks is only 50% of the total nucleon momentum. Hence the
other 50% of the nucleon momentum must be carried by partons which do
not partake in weak and electromagnetic interactions. Thus this points to the
presence of the strongly acting gluons as the other partons which carry half
of the momentum of the nucleon.
But the vexing issue of the constituent-quark versus the quasi-quark re-
mains. We shall demonstrate below that the difference between the quasi-
quarks and the constituent quarks will be made manifest by utilizing: (a) the
relativistic corrections to the non-relativistic quarks and thus using the con-
stituent quarks; and (b) the configuration mixed quark model wave functions
(in particular the deformed nucleon wave function).
Now assuming the quark model has no antiquarks and the nucleon contains
no s-quark,

Z1
F2ep (x) 1 1
Z Z
4 1
dx = u(x)dx + d(x)dx
x 9 0 9 0
0

4 1
= ×2+ ×1=1
9 9
and
232 Group Theory in Particle, Nuclear, and Hadron Physics

Z1 1 1
F2en (x)
Z Z
4 1
dx = d(x)dx + u(x)dx
x 9 0 9 0
0

4 1 2
= ×1+ ×2=
9 9 3
Now these results are completely independent of any details or differences
between any models, as these unpolarized structure functions simply weight
the quark numbers with the charges.
Thus as per these unpolarized structure functions one does not perceive
any difference between the quasi-quarks and the constituent-quarks. Both
yield identical results.
A pertinent question thus arises as to whether one may in fact ever witness
any physical evidence distinguishing quasi-quarks from constituent-quarks. In
fact, we find that the polarized structure functions are discriminating enough
to differentiate between the quasi-quarks and the constituent-quarks.
The unpolarized beams on unpolarized targets in inelastic electron scat-
tering therefore take the unpolarized structure functions as
1X 2 ↑
F 1 (x) = eq [q (x) + q̄ ↑ (x) + q ↓ (x) + q̄ ↓ (x)] (7.14)
2
Experiments have been completed with longitudinally polarized electron
and muon beams on longitudinally polarized hydrogen and deuterium targets.
These permit us to study the polarized structure functions:

1 X 2  ↑
q (x) + q −↑ (x) − q ↓ (x) + q −↓ (x)
 
g1 (x) = eq (7.15)
2
where eq is the quark charge (eq )e (e being the absolute magnitude of the
electronic charge). Here q ↑ and q ↓ correspond to the quarks with spin parallel
to, or antiparallel to, the proton spin or neutron spin. Thus we obtain polarized
structure functions g1p (x) and g1n (x) here.
An important sum rule relating the two polarized structure functions is
the Bjorken sum rule (derived by means of current algebra)
Z 1  
1 gA  αS 
(g1p (x) − g1n (x)) dx = 1− + ... (7.16)
0 6 gV π
where ggVA is the axial vector coupling constant obtained in n → p + e− + ν̄e
decay. Define
Z 1
g1p = g1p (x)dy (7.17)
0
Z 1
g1n = g1n (x)dx (7.18)
0
Quark Model 233

These are fully integrated spin-dependent structure functions.


These have been obtained experimentally, but the experimental determi-
nation of g1p in 1988 caused a major theoretical problem of proton physics,
which has come to be dubbed the “spin crisis”. We shall attempt to resolve
the troublesome “spin crisis” as being due to the basic difference between the
quasi-quarks and the constituent-quarks using symmetric principles. And as
discussed here, this shall be accomplished in terms of the quasi-quark picture
demanding corrections to the constituent-quark picture as: (i) relativistic cor-
rection; and (ii) corrections due to configuration mixing of wave functions in
the quark model (and in particular that of a deformed nucleon).
The “spin crisis” of proton studies was the fact that on the basis of
the quark model the integrated spin dependent structure function did not
match the experimentally determined value. The former theoretically calcu-
lated value is g1p = 185
= 0.278, while the EMC group [56] determined this
p
value to be g1 = 0.123. They did not obtain the value for the full range
0 ≤ x ≤ 1, but only for the interval of 0.01 < x < 0.7, which was 0.120.
Meanwhile the rest was obtained through smooth extrapolation.
Thus the observed fraction of spin of the proton is much less than what
one obtains from the quark model. We shall demonstrate below that the quark
model value of g1p = 18 5
= 0.278 demands that the entire spin of the nucleon
reside in quark spins only. Meanwhile, the above experimental value appears
to demand that the quarks carry only about 45% of the total spin of the
nucleon. Hence, clearly, the rest of the value of g1p has to arise from some
other unknown sources. This completely unexpected result was immediately
denominated the “spin-crisis” of proton science.
One may make two plausible assumptions that the SU(3)-flavour symmetry
holds and that the strange quark sea is unpolarized. Hence one may derive
separate sum rules for g1p and g1n as the Ellis-Jaffe sum rule:
 
gA !
Z 1 F
p(n) gV 5 3D −1
g1 (x)dx = ±1 + ( ) F (7.19)
0 12 3 D +1

where F and D are the anti-symmetric and symmetric coefficients of the SU(3)
symmetry. Using experimental values of D F
= 0.632 and ggVA = 1.254, we derive
p n
g1 = 0.189 and g1 = −0.002
Inconvenient incompatibility with the experimental result still persists.
Improved results by the E142 collaboration [57] are g1p = 0.129, and in
addition we have g1n = −0.031 (not determined earlier by the EMC group).
As if the original proton spin crisis was not enough, the new non-zero and
negative values of g1n obtained by the E142 collaboration creates an even
greater crisis for the quark model. As most of the models generically predict
g1n = 0, the E142 result is actually a complete calamity for several frameworks.
Combining the proton spin crisis and the neutron spin disaster, we term this
enlarged cataclysm the “spin conundrum”. To resolve this spin conundrum,
we shall investigate in more detail in Chapters 8 and 9.
234 Group Theory in Particle, Nuclear, and Hadron Physics

7.2 The Eightfold Way Model


We had assumed in the previous chapter that confinement in the quark
model arises from two possible mechanisms: the first one is from potential
models, which we shall elucidate in this chapter, and the second one is from
bag models, which we shall investigate in the next. In the former the quarks
are treated non-relativistically, while in the latter they are envisaged as rela-
tivistic.
While it is natural to view light quarks as relativistic entities, it is difficult
to justify visualizing them as non-relativistic objects. There does exist some
justification [58] for the latter approach, however. Here we accept this as a
phenomenological viewpoint. In fact this non-relativistic quark model had
preceded QCD by about a decade (1961 – 1964 vs. 1974).
Treating the (p,n) pair as the fundamental representation of the group
SU(2), the isospin group, was Heisenberg’s idea of a form of unification in
1933. As more and more baryons (besides the proton and the neutron) were
being discovered in the 1950s, the Sakata model, treating (p, n, Λ) as the
fundamental representation of the bigger SU(3) group, arose as an attempt to
bring some order in the understanding of hadrons.
The Sakata model was successful in accounting for the known 0− and 1+
mesons. However it was found to be wanting to account for the known spin 12
of a baryon then [30], [59]. Note that just as the (p,n) pair is fundamental in
the SU(2) framework, it is the (p, n, Λ) which is basic in SU(3) in the Sakata
model. In both systems all other hadrons (baryons and mesons) were presumed
to be built from these three.
Gell-Mann and independently Ne’eman, realized in 1961 that what was
correct in SU(2) was not holding good in the extrapolated SU(3). They re-
jected the fundamental three (p, n, Λ) of Sakata and suggested that it was the
number “eight” which was more basic. Inspired by the Irano-Scythian sage
Buddha Sakyamuni’s Eightfold Way Path [60] (eight paths, actually “middle
paths”, avoiding extreme luxury and extreme hardship) to attain Nirvana or
Salvation (the state of being free from suffering), he denominated his novel
idea as the eightfold way model [61].
The origin of the eightfold way model was the realization that there exists
+
a systematic parallelism between the 12 baryons and the 0− mesons, when
one supplements the SU(2) isospin number with a new quantum number called
the hypercharge Y. This is displayed in Table 7.1.
Now due to the new quantum number, the hypercharge Y, we have to go
beyond the confines of the group SU(2). If we take Y to be the generator of a
group U(1), then this extended group being SU (2)⊗U (1), can account for the
eight states given in Table 7.1. However the quantum number Y would not
be connected in any way with the isospin quantum number. But Gell-Mann
and Nishijima have suggested a physically well-satisfied connection (the well-
Quark Model 235
TABLE 7.1: Parallel structure
+
of 21 baryons and 0− mesons
leading to the eightfold way model
+
Y T 12 Baryon 0− Mesons
+1 12 p, n K +, K 0
0 1 Σ+ , Σ0 , Σ− π+ , π0 , π−
0 0 Λ0 η0
-1 21
Ξ , Ξ−
0 ¯
K , K−
0

known Gell-Mann-Nishijima relation) between T3 , the third generator of the


isospin SU(2) group, and Y in terms of the electric charge Q:
Y
Q = T3 + (7.20)
2
Here this expression connects independent generators of the group SU (2)⊗
U (1). However, if we demand that these two diagonal generators, T3 and Y,
belong to the same group, then this leads to the group SU(3).
In the eightfold way model fundamentality of the eight-dimensional repre-
sentation is demanded. If (p,n) forms a basic doublet (multiplet) of SU(2) than
+
it is required that in the eightfold way model the 0− and 21 representations
form a supermultiplet. So to say, we have a variety of compound quantum
numbers in the group SU(3) which is denominated “unitary spin”, which is
a generalization of the isospin that includes isospin and strangeness as one
single unit [7]. The eight particles (N, Σ, Λ, and Ξ) are in essence eight states
+
of just “one particle” with J p = 21 . No confusion exists in this conclusion,
for it is analogous to the visualization of (p,n) as two states of “one particle”,
the “nucleon”, in SU(2) isospin.
What should the situation be in the group SU(3)? Now there is a unique
representation of SU(3), the 32 − 1 = 8 dimensional adjoint representation. It
is most logical to associate the eight in the above eightfold way model with
the adjoint representation of SU(3). Let us study this adjoint representation
of these eightfold way model states given in Table 7.2.
To understand the adjoint representation in SU(3), first let us state that
the Cartan subalgebra for it is not λ3 , λ8 , the two commuting and diagonal
generators of SU(3). These form the Cartan subalgebra (as we shall see in the
next section) which will lead to the fundamental representation of SU(3). So
what is the Cartan Subalgebra of SU(3) for the adjoint representation?
We follow Cahn [62] and Dean [30] here. In the standard notation of all
the generators as already provided in Chapter 3, the SU(3) commutations are
all specified in Table 7.2.
Let there be vector space L with x, y ∈ L . Then we define the adjoint as
ad y(x) = [y, x] where [, ] is the commutator. With each x, y ∈ L pairs the
whole space and hence the dimension of ady is the dimension of the vector
236 Group Theory in Particle, Nuclear, and Hadron Physics
TABLE 7.2: The complete set of SU(3) commutation relations. The
columns are in the sequence: (T+ , T− , Tz , U+ , U− , V+ , V− , Y ). The same
sequence for the rows. The commutator is over the element from a row to
the one in a column.
0 2Tz −T+ V+ 0 0 −U− 0
−2Tz 0 T− 0 −V− V+ 0 0
T+ −T− 0 − 12 U+ 1
2 U−
1
2 V+ − 12 V− 0
1 3
−V+ 0 U
2 + 0 2 Y − Tz 0 T− −U+
0 V− − 12 U− − 32 Y + Tz 0 −T− 0 U−
0 −U+ − 12 V+ 0 T− 0 3
2 Y + Tz −V+
1
U− 0 2 V− −T− 0 − 23 Y − Tz 0 V−
0 0 0 U+ −U− V+ −V− 0

space. For SU(3) the adjoint space would be the size of the Lie algebra itself
and that is eight-dimensional.

————————————————–
Problem 7.1: Given [x, y] = z for a Lie algebra prove that the adjoint
representation preserves the commutation relation.
[adx, ady] = adz
————————————————–

Let us look at this eight-dimensional space L. Any operator adX can be


written as 8 × 8 matrices for a particular basis. Let us select the basis for L
as:
(V+ , U+ , T+ , Tz , T− , Y, U− , V− )
as in this order. The reason for this particular order shall become apparent in
a moment.

————————————————–
Problem 7.2: Determine the 8 × 8 matrices of T+ and T− for the above
basic vector.
————————————————–

With the above basis using Table 7.2 one obtains


Quark Model 237

1 
2

 − 12 


 1 

 0 
ad(Tz ) =   (7.21)

 −1 


 0 

 1 
2
− 12
and
 
1
 1 
 

 0 

 0 
ad(Y ) =   (7.22)

 0 


 0 

 −1 
−1
As ad (Tz ) and ad(Y ) are both diagonal, for any linear combination X =
aTz + bY , this ad(X) = ad(Tz ) + ad(Y ) is diagonal too,
1 
2a + b

 − 12 a + b 


 a 

 0 
ad(X) =    (7.23)
 −a 


 0 

1
2a − b
 
1
−2a − b

Thus (V+ , U+ , T+ , Tz , T− , Y, U− , V− ) are eigenvectors of ad(X) with eigen-


values 12 a + b, − 12 a + b, a, 0, −a, 0, 12 a + b, − 12 a + b , respectively.


Here we have basically determined the two-dimensional Cartan subalgebra


for the adjoint representation as given by X and Y as [X, Y ] = 0. As such each
of the elements of the above basis is ensured to be an eigenvector of X. It is
thus that we worked with the raising and lowering operators like T± , U± and
V± rather than Tx,y, Ux,y and Vx,y .
Note that for the adjoint representation the Cartan subalgebra is provided
by [X, Y ] = 0 when X = aTz + bY (and not [Tz , Y ] = 0). It may be il-
luminating to refer to the adjoint representation study of the SU(2) algebra.
Remember that physical pion states, as eigenstates of the charge operator, are
π + , π 0 , π − and not (π1 , π2 , π3 ). This requirement demands that we look at
the eigenstates of T± operators rather than Tx , Ty . So the Cartan subalgebra
for SU(2) is defined by the charge Q as an Abelian generator.
238 Group Theory in Particle, Nuclear, and Hadron Physics
1
The generalized situation is similar. In X above put a = 1 and b = 2 and
lo and behold, X = Q is the eigenvalue of Q̂ = adQ operator:
Y
Q = Tz + (7.24)
2
which is precisely the Gell-Mann-Nishijima expression for the electric charge
+
of the 12 baryons in the eightfold way model!
So as per our adjoint representation the chosen basis yields the charge
states of p, n, Σ+ , Σ0 , Σ− , Λ, Ξ0 , Ξ− , in this order exactly. Thus the eightfold

+
way model 21 baryon states are eigenstates of Q and Y operators. This is
demanded by the above study of the proper Cartan subalgebra of the adjoint
representation of SU(3).
It is important to note that it is incorrect to treat the adjoint representa-
tion, i.e., the octet of the eightfold way model, as an eigenstate of (T3 , Y ) but
it is actually an eigenstate of (Q, Y ). Note that in SU(2) the adjoint represen-
tation is an eigenstate of the charge operator Q. Also observe the consistency
that the subgroup structure of the adjoint representation of SU(2) within
SU(3) is being maintained here. So going from SU(2) to SU(3), as to the irre-
ducible representation of the adjoint state, one need only add Y to the charge
operator Q.
How is the eightfold way model description of the octet baryon as an
eigenstate of the electric charge Q and the hypercharge Y to be understood?
Let us return to SU (3) with a different Cartan subalgebra provided by isospin
T3 and hypercharge Y = √23 T8 . We define the 3-dimensional fundamental
representation as eigenstates of these two diagonal generators. Now singlets
and octets constructed from this in 3 ⊗ 3 ⊗ 3 = 1 + 8 + 8 + 10 require these
to be classified and distinguished by the SU (2)I Casimir operator I~2 [30].
Different eigenstates are distinguished as eigenstates I~2 . So these extrapolate
naturally to the isospin of the nucleus as (p,n) being eigenstates of nuclear
isospin SU (2)I . Now in contrast the above situation for the adjoint octet
model, there exists no isospin quantum number. And thus these adjoint states
are not defined by any isospin specified by any isospin quantum numbers.
This circumstance is perfectly natural as the adjoint representation is just
one single state and requires no Casimir operator to distinguish it from any
other state.
Now in the standard SU (2)I model of the nucleus the (p,n) pair has isospin
and it is this which is used in the Generalized Pauli Exclusion Principle to
define the various states in the nucleus. Most important an (n-p) pair being
made up of identical particles in the SU (2)I formalism is antisymmetric under
the interchange of its labels.
However as seen above the (p,n) pair constituted from the adjoint represen-
tation of the eightfold way model has no isospin quantum number to specify
itself. Thus this (n-p) pair is composed of distinguishable fermions and need
not be antisymmetric in their exchange. This is an amazing new prediction
Quark Model 239

of the eightfold way structure of spin 21 baryons. In short there is a duality


in the description of the spin 12 baryon octet as defined by the eightfold way
model and it is described as an octet in the SU (3)F model of quarks.
As we shall discuss in Chapter 11, this duality is justified and confirmed
by the study of nuclei as described simultaneously being composed of identical
fermions for the (n-p) pair in the SU (2)I model and independently as made
up of (n-p) treated as distinguishable particles.

7.3 SU (3)F Flavour Model


SU(2) and SU(3) groups were introduced in Chapter 3. We shall deal with
SU(2) in detail in Chapter 11. Here we shall consider SU(3) in detail, as it is
the most significant group in the quark model. We shall label it as SU (3)F to
signify the flavour structure, in contrast to another group SU (3)c character-
ized by the same group structure, but applied for the different physical reality
of colour. The latter endeavour results in the exact gauge theory of the strong
interaction, denominated Quantum Chromodynamics, which was elucidated
in the previous chapter. Each quark as a member of the group SU (3)F has
flavour (whether it is u−, d− or s− quark) and each has another independent
degree of freedom, namely colour (let us label it as R-, B- and G- colours; note
these are arbitrary labels but with time have become standardised by conven-
tion). First we use the SU (3)F group and then see as to how the requirement
of SU (3)c arises.
To avoid confusion, note that the group SU(n) for n = 2, 3 is a complete
mathematical structure as per the definition of the Lie groups in Chapter 3.
As such it is neutral and may be termed a “word” in this language of nature
(as per the definition of “word” in Appendix A). Now just as in ordinary
language we use the same “word” to refer to different entities (e.g., ‘bond’
describes stickiness, a share market bond or a legal bond, while ‘bumper’
denotes a bumper-crop or a bumper in an automobile). In the same manner
for a “word” of the mathematical language of nature, SU(2) is used to describe
the (n,p) isospin symmetry in a nucleus, or the spin degree of freedom of (↑, ↓)
for spin-1/2 entities, and for (h,t), the (helion,triton) degrees of freedom of
the nusospin group SU (2)A (see Chapter 11). SU(3) is here being utilized to
describe the flavour structure of quarks as well using it for the colour degree
of freedom. So the same word in group theory represents different physical
realities.
Now as per the SU (3)F group, the three flavours of u−, d− and s− quarks
are assumed to correspond to the fundamental representation of SU (3)F [28].
240 Group Theory in Particle, Nuclear, and Hadron Physics

 
u
¯ s̄

3 → d ; 3̄ = ū, d, (7.25)
s
Q = T3 + Y2 where T3 ≡ F3 = λ23 and Y = √23 F8 = √ λ8
3
.
Thus the eigenvalue of u−, d−, and s− quarks of the operator T3 , Y , and Q
are 12 , − 12 , 0 , 13 , − 13 , − 23 , and 23 , − 13 , − 13 , respectively. It is also assumed
 

to have baryon numbers of 31 .


As quarks possess fractional charges, but physically no known particle (as
detected in our physical world or in a laboratory) has fractional charges, so
these spin − 12 quarks should be confined inside physically detected hadrons
like protons, neutrons, pions, etc. So as to match the charge and the baryon
numbers, the baryons should be made of at least three quarks (an odd number)
and mesons of quark-antiquark pairs.
We know that for mesons in SU (3)F ,

⊗ = ⊕ (7.26)
3 ⊗ 3̄ = 1 + 8 (7.27)
For mesons thus one should obtain a singlet and an octet representation.
In spin space,

⊗ = ⊕
2⊗2=1⊕3 (7.28)
Thus spin 0 and spin 1. So a priori, on the basis of the SU (3)F model
the flavour state may exist in spin 0 and spin 1 states within the group space
SU (3)F ⊗ SU (2)S .
Now the baryons are constituted of three quarks. So we go from two to
three quarks as follows:

⊗ = ⊕
3 ⊗ 3 = 3̄ ⊕ 6 (7.29)

( ⊗ )⊗ =( ⊕ )⊗

=( ⊗ )⊕( ⊕ )
Quark Model 241

Finally,

⊗ ⊗ =( ⊕ )⊕( ⊕ ) (7.30)
The above in dimensional representation,

(3 × 3) ⊗ 3 = (3̄ + 6) ⊗ 3

= 3̄ ⊗ 3 ⊕ 6 ⊗ 3
Finally
3 × 3 ⊗ 3 = (1 ⊕ 8) ⊕ (8 ⊕ 10) (7.31)
Thus the SU (3)F model predicts that baryons should exist as a singlet, an
octet and a decuptet. But what are the two 8-dimensional octet states? The
difference between these, as is clear from the above discussion, is that the first
two particles in one are in a symmetric configuration and in the other one
these are in an antisymmetric configuration. For these there is no particular
symmetry with respect to the third quark. These are called “mixed symmetric
states”.
How about the spin of the states in the groups SU (3)F ⊗ SU (2)S ? In
SU (2)S group,

( ⊗ )⊗ =( ⊕ )⊗

=( ⊗ )⊕( ⊕ )
Finally,

⊗ ⊗ = ⊕( ⊕ ) (7.32)
The above in dimensional representation becomes

(2 ⊗ 2) ⊗ 2 = (1 + 3) ⊗ 2
=1⊗2⊕3⊗2
Finally,

2 ⊗ 2 ⊗ 2 = 2 ⊕ (2 ⊕ 4) (7.33)
Note that there is no Young Diagram with three boxes in a single column
for the group SU (2)F . So we obtain two mixed symmetric spin 21 states and
one symmetric spin 32 state.
242 Group Theory in Particle, Nuclear, and Hadron Physics

With this quark model we wish to explain the property of 0− and 1−


+ +
meson octet and 21 and 32 baryons octet and decuptet, respectively. These
are given in Figure 7.2 and Figure 7.3.
Let us now construct the quark wave function of mesons as per Equation
7.27 in the SU(3) space. Coupling states as per Appendix E are
X
|jm >= < j1 m1 j2 m2 |jm > |j1 m1 > |j2 m2 > (7.34)
m1 m2

We build π as π d¯ by matching the SU (3)F quantum numbers. As there is


+

no s-quark in the pion the SU (2)I isospin subgroup is relevant for it. Now the
¯ Similarly we find π − ∼ du.
CG coefficient 21 12 12 12 |11 = 1 and thus π + ∼ ud. ¯
0
For π we have
1 1 1 1 11 1 1
|I = 1, m = 0 >=< , , , − |10 > | >| − > (7.35)
2 2 2 2 22 2 2
1 111 1 1 1 1
+< − |10 > | − > | , > (7.36)
2 222 2 2 2 2
1
= √ −uū + dd¯

(7.37)
2
Note
 that in SU(2) the fundamental representation for quarks transforms
as ud while the conjugate representation for antiquark transforms as −u

d .

This is the source of the extra minus sign above. Had it been 2 ⊗ 2 then,
1
|I = 1, m = 0 >= √ (uu + dd) (7.38)
2
Next the Y=1 states K + , K 0 as doublets of isospin 12 are made up

 of
us̄ and ds̄, respectively. So also the Y = −1 isospin doublet of K̄ 0 , K̄ − as
¯ sū). Here all the SU (2) CG coefficients are the SU(2)- CG coefficients.
(sd,
(thus, e.g., T = 12 , Y = 1 (ds̄, us̄) isospin doublet). Out of nine (3 × 3̄) states
we have constructed seven meson states. Note that a particular invariant state
is the following singlet state
1 X 1
qi q¯i = √ uū + dd¯ + ss̄

Ψ1 = √ (7.39)
3 3

————————————————–
Problem 7.3: Show the state Ψ1 is an SU(3) invariant.
————————————————–

Now in spin 12 space, q, q combine to give spin 0 and spin 1. These thus give

0 and 1+ pseudoscalar and vector-octet mesons. Note that as per quantum
field theory the q, q state possesses an intrinsic parity of (-1). Thus the 0− and
1− mesons illustrated in Figure 7.2 have the wave functions as given. Note
Quark Model 243

+1 o +1
n P K K+

Y Y
P + P
1
J J 0
2
n+
+
Z z 0
n no
-1 0
+1 -i
*f n +]
1 o
3 1 0
+
1 3
" 2 2

0 0
K K

-1 -i
(a) (b)

Ko +1 K+

Y
p
I 1
0 +
P P P
-i 0 ( +i
1 U) o O
4
1 3
~2 2

0
K K

-i
(c)

FIGURE 7.2: Spin 1/2 baryons plus pseudoscalar and vector meson octets
244 Group Theory in Particle, Nuclear, and Hadron Physics

n + ++
+1

p +
3
J Y
2

zo
+*
z
-1 o I +:
_! + :i
2 2

_ * O

-2

FIGURE 7.3: Spin 3/2 baryon decuptet

¯ − 2s̄s is constructed to be orthogonal to



that the η wave function ūu + dd
the singlet state Ψ1 which is a scalar and associated with η 0 , which is the ninth
member of the q q̄ nonet.
As we shall demonstrate below this ninth member η 0 does not mix with the

0 pseudoselar octet, i.e., it is a genuine independent singlet from 3× 3̄ = 1+8.
However for the 1− vector mesons this singlet mixes with the octet and hence
the SU (3)F result 3 × 3 = 1 + 8 is not satisfied for the 1− vector mesons.
Consequently SU (3)F is a broken symmetry for 1− vector mesons but is a
good symmetry for 0− pseudo-scalar mesons. Why this is so is not explained
within the group SU (3)F ⊗SU (2)S itself. As we shall see below this particular
result requires us to go to a bigger SU (6)SF flavour group. SU (6)SF then
predicts the above difference between the 0− pseudo-scalar meson and the 1−
vector meson nonet structure. Thus this indicates that we should take the
group SU (6)SF as the better group to describe hadrons rather than the group
SU (3)F ⊗ SU (2)S . Still, for most pertinent calculations it is sufficient to stick
to the group SU (3)F ⊗ SU (2)S .
Next let us obtain the wave functions corresponding to the Young dia-
grams in Equation 7.29 and 7.30 in the groups SU (3)F ⊗ SU (2)S . First let us
construct spin states.
Now for two spin 12 states we get four states ↑↑, ↑↓, ↓↑, ↓↓. But these do not
correspond to the irreducible representation of the group SU (2). Or in other
words these are not states of “good” angular momenta of the group SU (2).
Quark Model 245

From quantum mechanics we know that for two spin 12 states, these should
be states of spin 0 and 1. From Young Diagrams in Equation 7.28, we see
that the first one holds for spin-0 and is antisymmetric, while the second one
stands for spin 1 and is symmetric. Let us define spinors as
1 1 1 −1 1 1
↑ ≡ χ 21 = | , >, ↓ ≡ χ 1 2 = | , − > (7.40)
2 2 2 2 2 2
The two spin states are,
1
|0, 0 >= √ (↑↓ − ↓↑)
2
|1, 1 >=↑↑
1
|1, 0 >= √ (↑↓ + ↓↑)
2
|1, −1 >=↓↓ (7.41)
The first one is a scalar and the other three are vector states. Now if we
bring in the third quark with spin 21 then as per Equation 7.32 and 7.33 we
have one symmetric state of dimension four and two two-dimensional mixed
symmetric states. The fully symmetric state with spin 23 is built by adding
the third spin to the vector part above and ensuring normalization to one as,
3 3
| , >↑↑↑
2 2
3 1 √
| , > [↑↑↓ + (↑↓ + ↓↑) ↑] / 3
2 2
3 1 √
| , − > [(↑↓ + ↓↑) ↓ + ↓↓↑] / 3
2 2
3 3
| , − >↓↓↓ (7.42)
2 2
These are the spin - 23 states of the decuptet baryons, symbolized by χS .
Now to the 10 totally symmetric decuptet states of the SU (3)F group.
With the fundamental representation we construct 10 states as,

uuu, ddd, sss, uud, uus, ddu, dds, ssu, ssd, usd (7.43)
Now the maximum stretched states like uuu (4++ ) , ddd (4− ) and
sss (Ω− ) are uniquely fully symmetric. Replacing ↑→ u, ↓→ d in Equation
7.42 we immediately obtain states for ∆+ and ∆− . Hence we use similar tech-
niques to obtain all the 10 states with consistent wave functions for the group
SU (3)F below. Let us call these Φs states.
246 Group Theory in Particle, Nuclear, and Hadron Physics

uuu √1 (duu + udu + uud) √1 (ddu + dud + udd) ddd


3 3
∆++ ∆+ ∆0 ∆−

√1 (suu + usu + uus) √1 (uds + dsu + sud + dus + usd + sdu)


3 6
+ 0
Σ ∗
Σ∗
√1 (sdd + dsd + dds)
3

Σ∗
√1 (ssu + sus + uss) √1 (ssd + sds + dss)
3 3
0 −

Ξ Ξ∗
sss
(7.44)
Ω−
Each of these have spin 32 called χS . Then the states in SU (3)F ⊗ SU (2)S
are given as ΦS χS ≡ (10, 4) with dimensionalities given for the two groups.
Thus the full wave function for the group SU (3)F ⊗ SU (2) is for example,
∆++ +
3 3 = (uuu) (↑↑↑) and for ∆ 3 1 =
1

3
(uud + udu + duu) √1
3
(↑↑↓ + ↑↓↑ + ↑↑↓).
2 2 2 2
∗0
Now forming an orthogonal state to the state Σ above gives us the totally
antisymmetric singlet states corresponding to the first Young Diagram on the
right in Equation 7.30
1
ΦA
singlet = √ (uds + dsu + sud − dus − usd − sdu) (7.45)
6
Note that the decuptet states are experimentally well determined, as are
the spin 21 octet states. But there is no empirically determined candidate
standing for the above singlet ΦA singlet state. So SU (3)F ⊗ SU (2)S group pre-
dicts one state too many. This is a puzzle here. However we shall see below as
to how the larger group SU (6)SF provides an unambiguous explanation for
why the singlet state should not figure in the quark model.
Next let us build the two mixed symmetric states of spin 21 defined by
these spin 12 states in Equation 7.30. In Equation 7.29 we built states of two
spin − 12 entities. From it, using the vector (spin-1) state we added one more
spin to construct the 32 states (Equation 7.42). Now using the scalar term in
Equation 7.41, where the first two spins are in an antisymmetric state, we add
one more spin to obtain
  √
χρ↑ = {(↑↓ − ↓↑) ↑} / 2
  √
χρ↓ = {(↑↓ − ↓↑) ↓} / 2 (7.46)

Here there is no specific symmetry with respect to the exchange of the


third spin. Another mixed symmetric state is obtained by using the vector
Quark Model 247
TABLE 7.3: Mixed symmetric wave functions for the spin-1/2
octet baryons
φλ ≡ φM,S φρ ≡ φM,A
1 1
P − √6 (ud + du)u − 2uud √ (ud − du)u
2
N + √16 (ud + du)d − 2ddu √1 (ud − du)d
2
Σ+ √1 (us + su)u − 2uus √1 (us − su)u
6 2
Σ0 √1 [s( du+ud
√ ) + ( dsu+usd
√ ) √1 [( dsu+usd
√ ) − s( ud+du
√ )]
6 2 2 2 2 2
du+ud
−2( √2 )s]
Σ− √1 [(ds + sd)d − 2dds
6
√1 (ds − sd)d]
2
s(du−ud)
Λ0 √1 [( dsu−usd
2

2
) + s( du−ud

2
)] √1 [
6

2
+ usd−dsu

2
− 2(du+ud)

2
s]
Ξ− √1 (ds + sd)s − 2ssd
6
√1 (ds − sd)s
2
Ξ0 − √16 (us + su)s − 2ssu √1 (us − su)s
2

part in Equation 7.41, where the first two spins are in symmetric states. By
adding another spin,

χλ↑ = − {(↑↓ + ↓↑) ↑ −2 ↑↑↓} / 6


χλ↓ = {(↑↓ + ↓↑) ↓ −2 ↓↓↑} / 6

(7.47)
Note that other notations found in the literature are χρ ≡ χM A and χλ ≡
χM S , indicating mixed symmetry with the first two states in antisymmetric
or symmetric states, respectively.
Now we supply the wave functions for the mixed symmetric states. We do
so in Table 7.3.
Note: For the corresponding spin wave functions denoted by χ replace u
and d by ↑ and ↓, respectively. So, for example, χρ− 1 = √12 (↑↓↓ − ↓↑↓),
2
The states φρ and φλ are called mixed symmetric states as an exchange
of spin third-state with, for example the spin first-state, mixes states with
definite symmetric properties under the exchange (1 ↔ 2) as
   √ 
χρ↑ = χρ↑ − 3χλ↑ /2 (7.48)
1↔3

————————————————–
Unsolved Problem 7.1: Check that the above expression is true.
————————————————–

By exchanging ↑→ u and ↓→ d we obtain the corresponding mixed sym-


metric states for the SU (2)I subset of SU (3)F . For all the states, we have to
build similar mixed symmetric states involving all the states as follows:

uud, uus, ddu, dds, ssu, ssd, uds, dsu (7.49)


248 Group Theory in Particle, Nuclear, and Hadron Physics

What we √ need to see is that


√ for the (u, s) and √ (s, d) subsets, one√ gets
(us + su) / 2 and (ds + sd) / 2 and (us − su) / 2 and (ds − sd) / 2 as
symmetric and antisymmetric states respectively. In addition the ss state is
symmetric too. These yield 6-dimensional symmetric and 3-dimensional anti-
symmetric states as discussed above. On top of these we build mixed symmet-
ric states which would be φρ (orφM A ) and φλ (orφM A ) states. These would be
8-dimensional, which we list in Table 7.3.
Now note that for the three quarks we have χρ , χλ states of SU (2)S and
φ , φλ state for SU (3)F groups. Remember that these ρ− and λ− states pos-
ρ

sess definite symmetries – symmetric or antisymmetric only for the exchange


of particle states number-1 and number-2, i.e., (1 ↔ 2). It turn out that for the
groups SU (3)F ⊗ SU (2)S for the states (8, 2) (i.e., 8-dimensional in SU (3)F
space and 2-dimensional SU (2)S space, whence spin 12 baryons). Now build
the following states
1
ΨS(8,2) = √ φρ χρ + φλ χλ

(7.50)
2
1
ΨA φλ χρ − φρ χλ

(8,2) = √ (7.51)
2
The first one is fully symmetric and the second one is fully antisymmetric
under all the exchanges of particles: 1 ↔ 2, 1 ↔ 3, 2 ↔ 3. It is important to
note that in the above the terms individually, i.e., φρ χρ or φλ χλ are symmetric
only on the exchange of 1 ↔ 2 with no specific symmetry with respect to the
exchange of the third particle. Only after having been added as in Equation
7.50 does full symmetry with respect to all the exchanges develop.

————————————————–
Unsolved Problem 7.2: Confirm the full symmetry and antisymmetry
of the ΦS(8,2) and ΦA
(8,2) states, respectively.
————————————————–

So given the two 8-states in 3⊗3⊗3 = 1⊕8⊕8⊕10 in SU (3)F for the group
SU (3)F ⊗SU (2)S we are able to obtain fully symmetric or fully antisymmetric
states as in Equations 7.50 and 7.51 respectively. A priori we would expect
+
the fully antisymmetric state like Equation 7.51 state, for the spin 21 octet
baryons, to work as per the Pauli Exclusion Principle for identical fermions.
Surprisingly, it fails and one finds that the symmetric wave function (Equation
7.50) works well to fit all the physically observable quantities for baryons.
First let us perform some simple calculations with this symmetric state wave
function in the quark model.
Quark Model 249

7.4 Quark Model Calculations

Mass Formulae

m (Σ) − m(N ) m(n) − m(p)


∼ 0.12 while ∼ 0.7 × 10−3
m (Σ) + m(N ) m(n) + m(p)
So we notice that though SU (2)I is a rather good symmetry, that of
SU (3)F is quite strongly broken. However assume that the binding energy
of quarks are independent of the quark flavour and the mass difference in par-
ticular of SU (3)F representation is entirely due to the quark mass difference
themselves and

mu = md = mū = md¯ = m1 and ms = ms̄ = m2 (7.52)



0 Pseudoscalar Mesons
P3
Meson mass is H = i=1 Hi (i=quark flavour)
Z

mk0 = (ψs̄ ψd ) H (ψs ψd ) dτ + m0
Z Z
= (ψs∗ Hs ψs dτ ) + (ψd∗ Hd ψd ) dτ + m0

= ms + md + mo = m1 + m2 + m0 (7.53)
Where mo is flavour independent common mass. Similarly.

mk¯0 = m1 + m2 + m0

mπ = 2m1 + m0
2 4
mη = m1 + m2 + m0 (7.54)
3 3
Thus
1
mK = (3mη + mπ ) (7.55)
4
This does not work that well (the left-hand term is 498 MeV while on the
right it is 446 MeV). However if squared masses are used m2k = 3m2η + m2π /4,
then it works pretty well (left term 0.25 Gev versus 0.23 Gev on the right
side). We as practical physicists use the square for the masses of mesons in
the quark model – and those do work much better than using mass terms
without squaring them. There has to be a fundamental reason as to why this
is so. We do not know for sure yet. An explanation commonly invoked is that
in the Lagrangian for mesons (as bosons) mass-squared terms arise while for
the baryons it is just plain mass terms.
250 Group Theory in Particle, Nuclear, and Hadron Physics

A somewhat better approach is to state that the 0− meson mass formula


should really be a formula for energies rather than masses mi . So replace mi
by Ei2 = m2i + p2i . Then

q p
3mη + mπ
q 3 m2η + p2η + m2π + pπ 2
mk = → m2k + p2k = (7.56)
4 4
Assume that in the limit of perfect symmetry the 0− mesons have zero
masses. For m → 0 we get
3pη + pπ
pk = (7.57)
4
But 0− mesons are not massless, while the mass arises due to some sym-
metry breaking. Let us use the relation m  p in the above,
 " ! #
m2k m2η m2π
 
pk 1 + = 3pη 1 + + pπ 1 + /4 (7.58)
2pk 2pη 2pπ
On using the above constraints on p’s only we obtain

3m2η + m2π
m2k = (7.59)
4
It worked!
1
Spin 2 Baryons

mN = 3m1 + B0
mΣ = 2m1 + m2 + B0
mΛ + 2m1 + m2 + B0
mΞ = m1 + 2m2 + B0 (7.60)
where B0 is some ground state flavour-independent common mass of baryons.
Thus we obtain:

mN + mΞ 3m∧ + mP
= (7.61)
2 4
This works well as the left side is 1127 MeV while the right side value of
1135 MeV.

————————————————–
3+
Problem 7.4: Obtain the mass formula for the baryon decuptet 2 . Then
show that the following equal spacing rule is observed,

mΩ− − mΞ∗ = mΞ∗ − mΣ∗ = mΣ∗ − m∆


Quark Model 251
+
(when the 32 baryon decuptet was predicted in 1964, the particle Ω− was
missing). This is how Ω was first predicted to exist with a clear-cut mass. It
was soon discovered in the laboratory and gave unequivocal support to the
concept of the SU (3)F group. Note the similarity with the Mendeleev table in
the late 19th century, with its predictions of missing chemical elements which
were subsequently discovered.
————————————————–

1− vector mesons
If 0− and 1− mesons octets are identical, then K, π, η and K octet would
correspond to K ∗ , ρ, ω and K¯∗ . Thus one predicts
3mω + mρ
mK ∗ = (7.62)
4
where on the left we obtain 892 MeV and on the right side we find 778 MeV.
This is a poor fit. We guess this may be due to ω(784), I = 0, S = 0 and
φ(1019), I = 0, S = 0 being close in mass, which could very well lead to their
mixture. Imagine a situation where SU(3) is an exact symmetry. Then φ8 and
φ1 would belong to the octet and the singlet states. Next we let this symmetry
be broken so as to produce physical φ and ω.

φ = φ8 cos θ + φ1 sin θ
ω = −φ8 sin θ + φ1 cos θ
where θ is the mixing angle. Given φ8 and φ1 as orthogonal in SU(3), then ω
and φ above are also orthogonal. Next assume ω is made up of u, ū, d, d¯ and
φ of ss̄. We derive:
1
ω = √ uū + dd¯

(7.63)
2
φ = ss̄ (7.64)
We know
1
φ8 = − √ uū + dd¯ − 2ss̄

6
1
φ1 = √ uū + dd¯ + ss̄

3
Therefore r r
1 1
ω=− φ8 + φ1
3 3
r r
2 1
φ= φ8 + φ1 (7.65)
3 3
q
1
Now with sin θ = 3 we obtain θ ∼ 350 .
252 Group Theory in Particle, Nuclear, and Hadron Physics

Here 1− is a true octet meson – ρ, φ8 , K 0∗ and thus they obey the 0−


octet formula,

3mφ8 + mρ = 4mk∗ (7.66)


Inverting from Equation 7.65,

2mφ + mω + mρ = 4mK ∗ (7.67)


Here on the left we obtain 3588 MeV and on the right side 3568 MeV,
which is a good agreement. Note no mass-squared terms are required here.

Electric Charge
Let us now calculate the charge of the proton. Given the charge operator
as
3
X
Q= Q(i) (7.68)
i=1

X3
↑ ↑ ↑
< P |Q| P >=< P Q(i) P ↑ >= 3 < P ↑ |Q(3)| P ↑ >


i=1

3
< φpρ |Q(3)| φpρ > + < φpλ |Q(3)| φpλ >

=
2
Now
1
< φpρ |Q(3)| φpρ >=< (udu − duu) |Q(3)| udu − duu >
2
 
1 2 2 2
= + =
2 3 3 3
Note we have used fixed position notation above. This means that for
the three particles the position of these particles is fixed in the left-to-right
sequence (123). So, for example, the first term above means that:

udu → u(1)d(2)u(3) (7.69)


3
and Q(3) acting on it picks up the 2 factor of the third position quark. Next,

1
< φpχ |Q(3)| φpχ >= < udu + duu − 2uud |Q(3)| udu + duu − 2uud >
6
 
1 2 2 4
= + − =0
6 3 3 3
So  
3 2
< P ↑ |Q| P ↑ >= +0 =1
2 3
Quark Model 253

————————————————–
Unsolved Problem 7.3: Show that the neutron charge equals zero in the
quark model.
————————————————–

Magnetic Moment
Next given that

σ+ | ↑>= 0, σ+ | ↓>= | ↑>, σz | ↑>= | ↑>,

σ− | ↑>= | ↓>, σ− | ↓>= 0, σz | ↓>= −| ↓>, (7.70)


Then

1 1 1
< χ↑ρ |σz (3)| χ↑ρ >= √ <↑↓↑ − ↓↑↑ |σz (3)| √ (↑↓↑ − ↓↑↑>= (1 + 1) = 1
2 2 2
(7.71)
↑ ↑ 1
< χλ |σz (3)| χλ >= <↑↓↑ + ↓↑↑ −2 ↑↑↓ |σz (3)| ↑↓↑ + ↓↑↑ −2 ↑↑↓>
6
1 1
= (1 + 1 − 4) = − (7.72)
6 3
1 1
< χ↑ρ |σ+ (3)| χ↑ρ >= √ <↑↓↑ − ↓↑↑ |σz (3)| √ (↑↓↑ − ↓↑↑) = 0 (7.73)
2 2
Let us define the magnetic moment operator as
3  
X e~
µz = σz (i) Q(i) (7.74)
i=1
2mq (i)c
So,

Q(i)e~ 2 e~ 1 e~
µq = , µu = , µd = − (7.75)
2mq (i)c 3 2mu c 3 2md c
The magnetic moment of proton is
3
X e~
µP =< P ↑ | σz (i)Q(i) |P ↑ > (7.76)
i=1
2m q (i)c

Q(3) ↑ e~
= 3 < P ↑ σ3 (i)

P >
mq (3) 2c
3 Q(3) p Q(3) p e~
= {< φpλ | |φ >< χ↑λ |σz (3)|χ↑λ > + < φpρ | |φρ >< χ↑ρ |σz (3)|χ↑ρ >}
2 mq (3) λ mq (3) 2c
(7.77)
Next,
254 Group Theory in Particle, Nuclear, and Hadron Physics


Q(3) p 1 Q(3)
< φpλ
mq (3) φλ >= 6 < udu + duu − 2uud mq (3) udu + duu − 2uud >

   
1 2 2 4 2 1 1
= + − = −
6 3mu 3mu 3md 9 mu md

Q(3) p 1 σ(3) 2
< φpρ

φ >= √ < udu − duu udu − duu >=
mq (3) ρ 2 mg (3) 3mu
So
      
3 2 1 1 1 2 e~ 8 1 e~
µP = − − + ×1 = +
2 9 mu md 3 3mu 2c 9mu 9md 2c
    
4 2 e~ 1 1 e~
= + − −
3 3 2muc 3 3 2mdc
4 1
µP = µu − µd (7.78)
3 3
We show in the problems that
4 1
µN = µd − µu (7.79)
3 3

—————————————-
Problem 7.5: Show that the neutron magnetic moment is µN = 43 µd −
1
3 µu .
————————————————–

If we take mu = md , then µu = −2µd


So
4 1
µp 3 (−2µd ) − 3 µd 3
= 4 1 =− (7.80)
3 µd − 3 (−2µd )
µN 2
This agrees very well with the experimental value (our prediction of -1.5
compared to the experimental value of -1.46). This stellar achievement was one
of the early successes of the symmetric quark model with the group structure
SU (3)F ⊗ SU (2)S
Next magnetic moment of ∆++ j= 32
Look at

∆++ = χs3 φs3 = | ↑↑↑> |uuu > (7.81)


2 2

3
X e~
< ∆++ |µ| ∆++ >=< ∆++ | σz (i)Q(i) |∆++ >
i=1
2mq (i)c
Quark Model 255


Q(3)
= 3 <↑↑↑ |σz (3)| ↑↑↑>< uuu
uuu > e~
mq (3) 2c
2 e~
= 3.1.
3mu 2c

µ∆++ = 3µu (7.82)

————————————————–
Problem 7.6: Show that the magnetic moment of Ξ− is - 31 µd + 43 µs .
————————————————–
Unsolved Problem 7.4: Show that the magnetic moment of the other
members of the baryon octet are

µΛ = µs (7.83)
4 1
µu − µs
µΣ+ = (7.84)
3 3
2 1
µΣ0 = (µu + µd ) − µs (7.85)
3 3
4 1
µΣ− = µd − µs (7.86)
3 3
1 4
µΞ0 = − µu + µs (7.87)
3 3
————————————————–

+
Thus these eight members of the spin 12 baryons using symmetric quark
model wave function have magnetic moments as given. Assuming mu = md
and using proton and Λ baryon magnetic moments as inputs, the other mag-
netic moments can be fitted as in Table 7.4. Also shown in Table 7.4 are the
experimental values.
The above shows how successfully the quark model fits all these magnetic
moments. Note that we have used the symmetric wave function here. What
happens if one utilizes a totally antisymmetric wave function in SU (3)F ⊗
SU (2) space? It turns out that these fail badly.
Below in the problem, the reader will first be asked to check that there
does exist an antisymmetric function in SU (3)F ⊗ SU (2)S space. And next
by calculating the magnetic moment of proton and neutron by using this
antisymmetric wave function, the dramatic failure of this enterprise shall be
emphasised.

————————————————–
Problem 7.7:
By explicitly constructing the full wave function verify that the state given
256 Group Theory in Particle, Nuclear, and Hadron Physics
TABLE 7.4: Magnetic moments of
spin-1/2 baryon octet in the quark model
Baryons SU (3)F -Theory Experiment
ρ 2.793 2.793
n -1.862 -1.913
Λ -0.614 -0.614
Σ+ 2.687 2.46
Σ0 0.825 -
Σ− -1.042 -1.16
Ξ0 -1.439 -1.25
Ξ− -0.508 -0.65


in Equation 7.51 ψA = (φλ χρ − φρ χχ ) / 2 is antisymmetric in the exchange
of any two states at position numbers 1, 2, and 3. Then demonstrate that
the antisymmetric states yield correct charges for the proton and neutron but
fail badly in matching the magnetic moments. Hence one cannot use such an
antisymmetric wave function of the three quarks in baryons within the group
SU (3)F ⊗ SU (2)S .
————————————————–

Weak Decays
Next we look at the weak decay of the neutron n → pe− ν̄e . At the quark
level, since p ∼ uud and n ∼ udd, this corresponds to the decay of d-quark
to u-quark. Spin-flip to non-spin-flip weak transition defines the axial vector
coupling constant as
P3
GA < p↑ | i=1 σ+ (i)τ+ (i)|n ↓>
gA = = P3 (7.88)
GV < p↑ | i=1 τ+ (i)|n ↑>
So we require τ+ |u >= 0, τ+ |d >= |u > and σ+ | ↑>= 0, σ+ | ↓>= | ↑>
etc.
We see that
X3
< p↑ σ+ (i)τ+ (i)|n ↓>
i=1
3n ↑ ↓ ↑ ↓
o
= < φpλ |τ+ (3)| φn p n
λ >< χλ |σ+ (3)| χλ > + < φρ |τ+ (3)| φρ >< χρ |σ+ (3)| χρ >
2
1 1
< φpρ |τ+ (3)| φnρ >=< √ (udu − duu) |τ+ (3)| √ (udd − dud) >= 1
2 2
1 1
< φpλ |τ+ (3)| φn
λ >=< − √ (udu + duu − 2udd) |τ+ (3)| √ (uud + dud − 2ddu) >
6 6
1 1
= − (1 + 1 + 0) = −
6 3
Quark Model 257

Next replacing p →↑ and n →↓ we get < χ↑λ |σ+ (3)| χ↓λ >= − 13 and
< χ↑ρ |σ+ (3)| χ↓ρ >= 1
So the numerator is = 32 − 13 × − 13 + 1 × 1 = 53
  

X 3
< p↑ τ+ (3) n↑ >= < φρρ |τ+ (3)| φnρ > + < φρλ |τ+ (3)| φnλ >

2
 
3 1
= 1− =1
2 3
So
5
gA = (7.89)
3
This is 1.667 while the experimental value is 1.25. This rather unsatisfac-
tory fit indicates that something is amiss here. In the next chapter we shall
show that a suitable D-state admixture (orbital angular momentum l=2) to
the ground states (for l=0) solves this awkward problem in the quark model.

————————————————–
Problem 7.8: Using the semi-leptonic decay Σ− → Σ0 e− ν¯e obtain the
axial vector coupling constant gA in the quark model.
————————————————–
Unsolved Problem 7.5: Demonstrate that for the semi-leptonic decay
Ξ− → Ξ0 e− ν̄e the axial vector coupling constant in the quark model is − 31 .


————————————————–

Hence the symmetric quark model (i.e., with totally symmetric wave func-
tion in the group space SU (3)F ⊗ SU (2)S ) does pretty well in fitting the
masses of the hadrons, performs very well for the magnetic moments of all the
+
baryons of the 21 octet, and does acceptably well for the axial vector cou-
pling constants. Meanwhile, the antisymmetric wave function for the group
SU (3)F ⊗ SU (2)S fails miserably.
So how come baryons, as fermionic objects made up of 3 quarks (an odd
number) in the degrees of freedom available in SU (3)F ⊗SU (2)S , are failing to
produce the fermionic character of the quark model? As quarks exist in our 3-
dimensional configuration space, is it possible to bring out its fermionic nature
by including the orbital symmetric structure? We expect that for the three
quarks the relative orbital angular momentum for the ground stage should be
l=0. Indeed it is possible to construct antisymmetric states of 3 quarks in the
relative angular momentum l=0 state [63].
Let the completely antisymmetric wave function in the orbital space be
defined as f (r~1 , r~1 , r~1 ) where quarks sit as positions r~1 , r~2 , and r~3 , respectively.
Then one can define the baryonic charge density as
258 Group Theory in Particle, Nuclear, and Hadron Physics

Z
2
ρ (~r) = d3~r |f (r~1 , r~2 ) − (r~1 + r~2 )| (7.90)

Above we have chosen the coordinates of the 3-quarks in the centre-of-mass


system in which

r~1 + r~2 + r~3 = 0 (7.91)


As function ‘f’ is antisymmetric the above integral vanishes and thus the
charge density of baryons is zero in this particular case. So as per this anti-
symmetric state there should be a “hole” at the centre of a nucleon. But this
is in contrast to the experimentally determined charge density of the baryons.
Thus this wave function cannot be taken seriously in the quark model.
Another problem in using such an antisymmetric orbital wave function for
3-quarks in the l=0 state was aptly pointed out by Mitra and Majumdar [64]
They emphasised that with this l=0 wave function one would observe zeroes
in the form factor of the proton. Experimentally, no such zeroes have been
detected. Hence this also rules out the l=0 antisymmetric wave function.
Thus with the degrees of freedom available in the group SU (3)F ⊗SU (2)S ⊗
SO(3)l (where the last one is for the 3-dimensional orbital space) it is not phys-
ically possible to obtain a totally antisymmetric wave function for baryons.
Now this is a very puzzling enigma. In the 1960s the crisis was so severe that
many a scientist was even willing to forgo the Pauli Exclusion Principle for
baryons. They claimed that the Pauli Exclusion Principle holds good for two-
protons and two-electrons, but in the case of three fundamental entities as in
the quark model it may very well be violated.
Another approach taken was to resort to additional degrees of freedom.
For it was apparent that not enough parameters were available within the
group structure SU (3)F ⊗ SU (2)S ⊗ SO(3)l to obtain a proper antisymmetric
function. So to say, this group structure may have just enough freedom to
provide us with a symmetric function, but not with an antisymmetric one.
Consequently it was hypothesized that invoking an additional unknown de-
gree of freedom for quarks could perhaps permit the retention of the otherwise
successful Pauli Exclusion Principle. By the late 1960s this additional degree
of freedom had come to be called colour. In this novel extension, one simply
assumes that each quark flavor (u,d,s) possesses this new additional degree of
freedom. A minimum of three colours, say R, B, G (red, blue, green, respec-
tively) are required to obtain an antisymmetric wave function for the three
quark baryons. Now the total wave function of baryons in the quarks model
is


Ψbaryon = ψSU 3)F · χSU (2)S · φSO(3)l Symmetric · {ℵColour }Antisymmetric
(7.92)
where in SU (3)F ⊗ SU (2)S ⊗ SO(3)l the wave function is fully symmetric and
the colour space wave function is antisymmetric as
Quark Model 259

1
{ℵColour }Antisymmetric = √ (RGB + GBR + BRG − RBG − BGR − GRB)
6
(7.93)
Thus the total wave function Ψ is a fully antisymmetric wave function.
This happily explains the success of, and eminently justifies the use of, the
symmetric quark model wave functions.
If the colour degree of the freedom had no role to play in physics other than
to just arbitrarily fix the above symmetry problem, then its role may clearly be
debated. However it turned out that this lucky guess actually manifests itself
more explicitly and directly in hadron physics. Actually three colours is now
an empirically determined fact. In fact the group associated with it is SU (3)c ,
and which leads to an exact gauge theory called Quantum Chromodynamics
(which was studied in Chapter 6) and forms a very successful theory of the
strong interaction.

7.5 SU (6)SF Model


Wigner had joined the SU (2)I isospin of np as a fundamental rep-


resentation and the SU (2)S of spin to form a bigger group SU (4)IS ⊃


SU (2)I ⊗ SU (2)S . One assigns the fundamental representation of SU (4)IS
as
 ↑
p
 p↓ 
 ↑
n  (7.94)

n
If the forces between nucleons in the nucleus are independent of the spin
and the isospin degrees of freedom, then SU (4)SI may be a reasonable ap-
proximate symmetry of the nucleus. However, at high energies, as spin and
orbital angular momentum mix, hence SU (4)SI may at best be a good sym-
metry at low non-relativistic velocities. Hence the isospin SU (2)I is a good
approximate symmetry at any energy, but SU (4)SI may be a good dynamical
symmetry only at low energies and non-relativistic velocities.
Note that this SU (4)SI is different from the SU (4)F used for four-flavour
of quarks (u, d, s, c), where it stands as its fundamental representation. One
can build baryons with charm c-quark by including these in a product 4 ⊗ 4⊗
and 4 ⊗ 4 ⊗ 4̄ states.
260 Group Theory in Particle, Nuclear, and Hadron Physics

————————————————–
Unsolved Problem 7.6: Using the Young Diagram technique determine
the irreducible representations in the product 4 ⊗ 4 ⊗ 4 and 4 ⊗ 4̄.
————————————————–

As of now we know of six independent flavours of quarks called u, d, s, c, b,


t (c:charm, b:beauty, t:top quark). One takes these to form the fundamental
representation of the largest flavour group SU (6)F as the simplest generaliza-
tion of the group SU (3)F . Baryons in this group are taken as belonging to the
irreducible representation of the products 6 ⊗ 6 ⊗ 6 and the mesons of 6 ⊗ 6.
We shall study this group SU (4)IS in Chapter 11. Using this analogy,
Guersey and Radicati and Sakita combined the groups SU (3)F and SU (2)S
[30],[42] into a larger group SU (6)SF ⊃ SU (3)F ⊗ SU (2)S where the funda-
mental representation of this group is given by the vector:
 ↑
u
u↓ 
 ↑
d 
 ↓
d  (7.95)
 ↑
s 
s↓
Here u↑ connotes u-quarks with spin ↑ and so forth. We take the 6 and 6̄
fundamental representation reduced with respect to the subgroups SU (6)SF ⊃
SU (3)F ⊗ SU (2)S as
 
6SU (6)SF → 3SU (3)F , 2SU (2)S
 
6̄SU (6)SF → 3̄SU (3)F , 2SU (2)S (7.96)
Note that this SU (6)SF group is completely different from the group
SU (6)F , relevant to the six-flavours of quarks, as defined above.
One can expect the SU (6)SF symmetry to be a dynamical symmetry to
be valid only at low energies. In SU (6)SF flavour and spin become indistin-
guishable and can be transformed into each other. A Lorentz boost on SU (6)
will mix spin and orbital angular momentum. But SU (3)F as an internal
symmetry, will not mix with the orbital angular momentum. Thus a Lorentz
boost breaks the SU (6)SF symmetry by distinguishing between flavour and
spin. So SU (6)SF may work only at low energies for non-relativistic velocities.
Also spin and orbital angular momentum should conserve separately for an
interaction which is invariant under the group SU (6)SF .
Let us build states for the baryons and the mesons in the group SU (6)SF
Baryons = qqq :

⊗ ⊗ =( ⊕ )⊕( ⊕ ) (7.97)
Quark Model 261

6 ⊗ 6 ⊗ 6 = 20A ⊕ 70M A ⊕ 70M S ⊕ 56S (7.98)


Meson: qq :

⊗ = ⊕ (7.99)

6 × 6 = 1 + 35 (7.100)
Next let us see the flavour and the spin structures of these irreducible
representations by decomposing with respect to the subgroups SU (6)SF ⊃
SU (3)F ⊗ SU (2)S . (Note no diagram with three boxes in a single column for
the SU (2) group). For the fully symmetric state we have

S →( F , S )⊕( F , S ) (7.101)

56 → (10F , 4S ) ⊕ (8F , 2S ) (7.102)


So the symmetry has flavor and the spin structure given as

56S = (10, 4) ⊕ (8, 2) (7.103)


Now the virtue of the state 56S is evident for all to see. Only octet and
decuptet baryons with correct spins exist. There is no singlet present here
(as it was there in SU (3)F model). So the puzzle of the SU (3)F model, with
regards to the presence of the singlet is finally resolved here. The 56S state
in the SU (6)SF model has no singlet state. This indicates the colossal power
and deep-rooted significance of the product group SU (6)F S .
The elegant decomposition of the other states in the above product (as
these may be relevant for the excited states of the baryons) yields

A →( F , S )⊕( F , S ) (7.104)

20A → (8F , 2S ) ⊕ (1F , 4S ) (7.105)


Similarly one can show that

70 = (10, 2) ⊕ (8, 4) ⊕ (8, 2) ⊕ (1, 2) (7.106)


262 Group Theory in Particle, Nuclear, and Hadron Physics

It turns out that the states Equation 7.105 and 7.106 do not yield sat-
isfactory descriptions of the ground states of baryons. Meanwhile, Equation
7.102 with (8, 2) and (10, 4) in SU (3)F ⊗ SU (2)S group describes the physical
reality of the baryon well, as depicted in Figure 7.2 and Figure 7.3. Hence we
represent
1
(8, 2) = √ (φ% χ% + φλ χλ ) (7.107)
2
(10, 4) = (φs χs ) (7.108)
where φρ ≡ φM A and φλ = φM S are given Table 7.3. This is exactly the
wave function we had utilized above in the quark model calculations for the
group SU (3)F ⊗ SU (2)S . But now the difference is that in Equation 7.107
it is a decomposition, while in Equation 7.50 it corresponds to the group
products as taken therein. Thus the states make up u↑ , d↓ and so forth of
SU (6)SF as elementary entities, rather then the product states u⊗ ↑ and
d⊗ ↓, respectively, of SU (3)F ⊗ SU (2)S groups.
To make it clear we derive below the SU (6)SF form of the baryon octet
and baryon decuptet states. So for spin-3/2 delta,

1 1
∆+
↑ = φs χs = √ (uud + udu + duu) √ (↑↑↓ + ↑↓↑ + ↓↑↑) (7.109)
3 2
1
= √ (u↑ u↑ d↓ + u↑ u↓ d↑ + u↓ u↑ d↑ + u↑ d↑ u↓ + u↑ d↓ u↑
3
+u↓ d↑ u↑ + d↑ u↑ u↓ + d↑ u↓ u↑ + d↓ u↑ u↑ ) (7.110)
This is the complete SU (6)SF wave function of ∆+ ↑ . Note that the first
one is the SU (3)F ⊗ SU (2)S decomposition of the state and the last one is
the full SU (6)SF state. What is being done is that the product state like
 ↑ (1) of SU (3)F ⊗ SU (2)S , both sitting at position ‘1’, is taken to
u(1)⊗
u↑ (1) of SU (6)SF as a single state of a u-quark with an internal spin ↑,
as a single entity sitting at position ‘1’. It is a single state elementary wave
function rather than a composite as in SU (3)F ⊗ SU (2)S .
Next for proton,
1  
|p↑ >= √ φpρ χ↑ρ + φpλ χ↑λ (7.111)
2
1 −1 −1
= √ [ √ (udu + duu − 2uud) √ (↑↓↑ + ↓↑↑ −2 ↑↑↓)
2 6 6
1 1
+ √ (udu − duu) √ (↑↓↑ − ↓↑↑)]
2 2
 
1 1 1
= √ [udu (↑↓↑ + ↓↑↑ −2 ↑↑↓) + (↑↓↑ − ↓↑↑)
2 6 2
Quark Model 263
 
1 1
+duu (↑↓↑ + ↓↑↑ −2 ↑↑↓) − (↑↓↑ − ↓↑↑) ]
6 2
 
1
+uud − (↑↓↑ + ↓↑↑ −2 ↑↑↓)
3
   
1 2 1 1 1 2 1
= √ [udu ↑↓↑ − ↓↑↑ − ↑↑↓ + duu − ↑↓↑ + ↓↑↑ − ↑↑↓
6 3 3 3 3 3 3
 
1 1 2
−uud ↑↓↑ + ↓↑↑ − ↑↑↓ ]
3 3 3
So,

1
|p↑ >= √ [2u↑ d↓ u↑ − u↓ d↑ u↑ − u↑ d↑ u↓ − d↑ u↓ u↑ + 2d↓ u↑ u↑
18

−d↑ u↑ u↓ − u↑ u↓ d↑ − u↓ u↑ d↑ + 2u↑ u↑ d↓ ] (7.112)


This is the complete SU (6)SF wave function for the proton with spin-up,
p↑ .

————————————————–
Problem 7.9: Obtain the SU (6)SF wave function of a neutron with spin
up, n↑ (b) Σ+ spin up, Σ+↑ .
————————————————–

We depict the pseudoscalar and vector meson octet in Figure 7.2. Let us
decompose SU (6)SF ⊃ SU (3)F ⊗ SU (2)S for mesons. Take 6 ≡ (3, 2) and
6̄ = (3̄, 2) to be broken up as per the above subgroups. Then

6 ⊗ 6̄ = (3, 2) ⊗ (3̄, 2)
(3 ⊗ 3̄, 2 ⊗ 2) = (1 ⊕ 8, 1 ⊕ 3)
6 ⊗ 6̄ = (1, 1) ⊕ (8, 1) + (8, 3) ⊕ (1, 3)
The first term is singlet with spin-0, the second one octet with spin-0, the
third one octet with spin 1 and the last one is singlet with spin-1. Note that
the singlet of SU (3)F ⊗ SU (2)S must go to the singlet of SU (6)SF . So,

6 ⊗ 6̄ = 1 ⊕ {(8, 1) ⊕ (8, 3) ⊕ (1, 3)} (7.113)


6 ⊗ 6̄ = 1 ⊕ 35 (7.114)
where
35 = {(8, 1) ⊕ (8, 3) ⊕ (1, 3)} (7.115)
The 35 representation in SU (6)F S is an irreducible representation and as
broken up above, tells us that for spin-1 the octet and the singlet join together
264 Group Theory in Particle, Nuclear, and Hadron Physics

and thus form a nonet. As we saw earlier, however, as the spin 0− octet belongs
to the 35 representations and the 0− singlet is outside it, these do not mix.
Thus this explains the lack of any significant mixing of the singlet and the
octet for the 0− mesons. So what was a puzzle for the quark model with the
group structure SU (3)F ⊗ SU (2)S is resolved fundamentally in the SU (6)SF
group.

————————————————–
Problem 7.10:
We know that SU (4)F is the group to describe the four flavour u-,d-,s-,c
quarks. Take SU (8)SF ⊃ SU (4)F ⊗ SU (2)S . Just as in the SU (6)SF case,
using Young Diagrams, obtain the product of baryons (qqq) and mesons (qq)
in the SU (8)SF group to derive their reduction with respect to the above
subgroup structure.
————————————————–
Problem 7.11: Given three flavours – (u, d, s), we wish to study these
as ((u, d) + s) to be able to obtain states with s-quark content separated
out. Thus we take SU (6) ⊃ SU (4) ⊕ SU (2) where 6 = 4 + 2. In this the
fundamental representation of SU (6) is split into the SU (4) ⊃ SU (2)I ⊗
SU (2)S fundamental representation and the 2- in SU(2) has the fundamental
↑ 
representation of ss↓ . Find all the irreducible representations of 6 ⊗ 6 ⊗ 6
and 6 × 6 under this group decomposition.
————————————————–

7.6 Solutions of Problems

Solution 7.1:

[adx, ady] v = [x, [y, w]] − [y, [x, w]]


= [x, [y, w]] + [y, [w, x]]
= − [w, [x, y]]
= [[x, y] , w] = [z, w] = adz(w)
————————————————–
Quark Model 265

Solution 7.2:
 
0 0 0 1 0 0 0 0
0 0 0 0 0 0 0 0
 
0 0 −1 0 0 0 0 0
 
0 2 0 0 0 0 0 0
ad T+ = 
0 0 0 0 0 0 0

 0

0 0 0 0 0 0 0 0
 
0 0 0 0 0 0 −1 0
0 0 0 0 0 0 0 0
————————————————–
Solutio 7.3:


X X X
q̄j qk Uki U −1

q¯i qi = q̄j Uji qk Uki = ij
i ijk ijk
X X
= q̄j qk δjk = q̄i qi
jk

Hence this singlet states is invariant under the SU(3) transformation.


————————————————–
Solution 7.4:

m∆ = 3m1 + B1
mΣ∗ = 2m1 + m2 + B1
mΞ? = m1 + 2m2 + B1
mΩ = 3m2 + B1
Equal spacing rule is

mΩ− − mΞ∗ = mΞ∗ − mΣ∗ = mΣ∗ − m∆ = m2 − m1


It is well-observed empirically.
————————————————–
Solution 7.5:

3 N Q(3) N ↑ ↑ N Q(3) N ↑ ↑ e~
µN = {< φλ | |φ >< χλ |σz (3)|χλ > + < φρ | |φ >< χρ |σz (3)|χρ >}
2 mq (3) χ mq (3) ρ 2c

Q(3) N
< φN
ρ
φ >= − 1
mq (3) ρ 3md
 
N Q(3) N 1 1 4

< φλ φ >= − +
mq (3) λ 9 md mu
266 Group Theory in Particle, Nuclear, and Hadron Physics

So
      
3 1 1 4 1 1 e~
µN = − + − + − ×1
2 9 md mu 3 3md 2c
 
2 2 1 e~
=− +
9 md mu 2c
   
4 1 e~ 1 2 e~
= − −
3 3 2md c 3 3 2mu c
4 1
µd − µu µN = (7.116)
3 3
————————————————–
Solution 7.6:


X
− − e~ −
< Ξ |µz | Ξ >=< Ξ
σz (3)Q(3) Ξ >
2mq (3)c
 
3 Ξ− Q(3) Ξ− ↑ ↑ Ξ− Q(3) Ξ− ↑ ↑
= < φρ m (3) φp >< χρ |σz (3)| χρ > + < φλ m (3) φλ >< χλ |σz (3)| χλ >

2 q q

Ξ− Q(3) Ξ− Ξ− Q(3) 1 1

< φρ φp >=< φρ √ (dss − sds) >= −
mq (3) mq (3) 2 3ms
− Q(3) Ξ− − Q(3) 1 1 1 2
< φΞ
λ | |φλ >=< φΞ
λ | | − √ (dss + sds − 2ssd) >= − ( + )
mq (3) mq (3) 6 9 ms md
So  
1 4 e~
< Ξ− |µz | Ξ− >= −
9md 9ms 2c
     
1 1 e~ 4 1 e~
= − − − − −
3 3 2md c 3 3 2ms c
1 4
µΞ− = − µd + µs
3 3
————————————————–
Solution 7.7:

3n o
< P ↑ |Q| P ↑ >= < χ↑λ |χ↑λ >< φpρ |Q(3)| φpρ > + < χ↑ρ |χ↑ρ >< φpλ |Q(3)| φpλ >
2
 
3 2
= +0 =1
2 3
Similarly
< N ↑ |Q| N ↑ >= 0

3
X e~
µP =< P ↑ | σz(i) Q(i) |P ↑ >
i=1
2mq (i)c
Quark Model 267
 
3 ↑ ↑ p Q(3) φp > + < χ↑ |σz (3)| χ↑ >< φp Q(3) φp >

= < χλ |σz (3)| χλ >< φρ ρ ρ ρ λ λ
2 2m (3)c
q

q2m (3)c
    
3 1 2 2 1 1 e~
= − +1× −
2 3 3mu 9 mu md 2c
1
− e~
 
3 2 2 1 2 1 e~
= − + − = 3 = µd
2 9mu 9 mu 9 md 2c 2md c
     
3 1 1 4 1 1 e~ 2 e~
µN = 1× − + + − − = = µu
2 9 md mu 3 3md 2c 3 2mu c
Now,
µp µd µP 1
= T aking µu → −2µd , get =−
µN µu µN 2
which fails as experimentally µµN
P
∼ − 32
Hence such an antisymmetric wave function for the proton and neutron
is rejected. However as we saw above a symmetric wave function succeeds in
obtaining these magnetic momenta correctly and therefore these are accepted
in the quark model.
————————————————–
Solution 7.8:
P
3
< Σ0↑ i=1 σ+ (i)τ+ (i) Σ−↓ >

ga = P
3

< Σ0↑ i=1 τ+ (i) Σ−↑ >


X 3
< Σ0↑ σ+ (i)τ+ (i) Σ−↓ >


i=1
 
3 Σ0 Σ− ↑ ↓ Σ0 Σ− ↑ ↓
= < φρ |τ+ (3)| φρ >< χρ |σ+ (3)| χρ > + < φλ |τ+ (3)| φλ >< χλ |σ+ (3)| χλ >
2

Σ0 Σ− 1 1 1
< φρ |τ+ (3)| φρ >=< (dsu − sdu + usd − sud) |τ+ (3)| √ (dsd − sdd) >= √
2 2 2
0
Σ− 1
< φΣλ |τ+ (3)|φλ >=< √ (usd + sud + dsu + sdu − 2uds − 2dus)
2
1
|τ+ (3)| √ (dsd + sdd − 2dds) >
6
1
=< · · · | √ (dsu + sdu + 0) >
6
1
≡ √
3 2
n o 2√2
So numerator = 32 √12 × 1 + 3√ 1
2
− 1
3 = 3
268 Group Theory in Particle, Nuclear, and Hadron Physics

and
3 o √
X 3n 0
Σ− 0
Σ−
< Σ0↑ | τ+ (i)|Σ−↑ >= < φΣ
ρ |τ+ (3)| φρ > + < φΣ
λ |τ+ (3)| φλ > = 2
i=1
2

 2
gA Σ− → Σ0 e− ν̄e =

3
————————————————–
Solution 7.9:
(a)
1  
|n↑ >= √ φnρ χ↑ρ + φnλ χ↑λ
2
1 1 −1
= √ [ √ (udd + dud − 2ddu) × √ (↑↓↑ + ↓↑↑) − 2 ↑↑↓
2 6 6
1 1
+ √ (udd − dud) √ (↑↓↑ − ↓↑↑]
2 2
1
|n↑ >= √ [−2d↑ u↓ d↑ − 2d↑ d↑ u↓ − 2u↓ d↑ d↑ + d↑ d↓ u↑ + u↑ d↓ d↑ + d↑ u↑ d↓
18
+u↑ d↑ d↓ + d↓ d↑ u↑ + d↓ u↑ d↓ ]
(b)
1  + ↑ Σ+ ↑

|Σ+↑ >= √ φΣ ρ χ ρ + φ λ χλ
2
1
|Σ+↑ >= √ [2u↑ u↑ s↓ + 2s↓ u↑ u↑ + 2u↑ s↓ u↑ − u↑ u↓ s↑ − s↑ u↓ u↑ − u↑ s↑ u↓
18
−u↓ u↑ s↑ − s↑ u↑ u↓ − u↓ s↑ u↑ ]
————————————————–
Solution 7.10:
(qqq) for baryons:

⊗ ⊗ =( ⊕ )⊕( ⊕ )

8 ⊗ 8 ⊗ 8 = 56A ⊕ 168M A ⊕ 168M S ⊕ 120S


Note that the 56-dimensional representation here is antisymmetric, in con-
trast to the SU (6)SF case where the 56-dimensional representation is symmet-
ric. SU (4)F ⊗ SU (2)S content of SU (8)SF is,

120S ≡ (20s , 4) ⊕ (20M S , 2)


Quark Model 269

168M S = (20S , 2M S ) ⊕ 4, 2M S ⊕ (20M S , 4M S ) ⊕ (20M S , 2M S )
168M A = (20S , 2M A ) ⊕ (4̄M S , 2M A ) ⊕ (20M A , 4S ) ⊕ (20M A , 2M A )
56A = (4¯A , 4) ⊕ (20M A , 2))
Mesons as q q̄:

⊗ = ⊕

8 ⊗ 8 = 1 ⊕ 63
63 ≡ (15, 3) ⊕ (15, 1) ⊕ (1, 3)
and singlet is the same in SU (8)SF and SU (4)F ⊗ SU (2)S .
————————————————–
Solution 7.11:
In SU (6), under the decomposition SU (6) ⊃ SU (4) ⊗ SU (2). Take ( • )
as the singlet state.

=( , •)⊕(•, )

6 = (4, 1) ⊕ (1, 2)

=( , •)⊕(•, )

6̄ = (4̄, 1) + (1, 2)
Then one finds

=( , • )⊕( , ) ⊕( , ) ⊕( • , )

56 = (20, 1) ⊕ (10, 2) ⊕ (4, 3) ⊕ (1, 4)


270 Group Theory in Particle, Nuclear, and Hadron Physics

=( , •)⊕( , ) ⊕( , )


20 = 4, 1 ⊕ (6, 2) ⊕ (4, 1)

N ote that in SU (2) • is f or and does not exist.


We also get

70 = (200 , 1) ⊕ (10, 2) ⊕ (6, 2) ⊕ (4, 1) ⊕ (4, 3) ⊕ (2, 1)

Next for mesons:



35 = (15, 1) ⊕ (1, 1) ⊕ 4, 2 ⊕ (4, 2) ⊕ (1, 3)
————————————————–
Chapter 8
Bag Models

8.1 Why a Bag? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271


8.2 Confinement in a Spherically Static Bag . . . . . . . . . . . . . . . . . . . . . . . . 272
8.3 MIT Bag Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275
8.4 Finite Mass Quarks in a Bag . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 292
8.4.1 Magnetic Moments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295
8.4.2 Axial Vector Coupling Constant . . . . . . . . . . . . . . . . . . . . . . . . 296
8.4.3 Spin Structure of the Nucleon . . . . . . . . . . . . . . . . . . . . . . . . . . . 297
8.5 Scalar and Vector Confining Potentials . . . . . . . . . . . . . . . . . . . . . . . . . . 300
8.6 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303

8.1 Why a Bag?


In the previous chapter we treated the u-, d-, and s-quarks as non-
relativistic in nature, with masses of the order of 300 MeV for the up and
down quarks, and about 500 MeV for the strange quarks. This works very
well in the phenomenological quark model.
However, experimentally we know that quarks inside hadrons behave as
if they are asymptotically free. This implies that these individual quarks at
high energies and high momentum transfers behave as if they are weakly
interacting. They are relativistic and have zero or very small masses. These
are current quark masses in contrast to the constituent quark masses of the
non-relativistic quark model. How can we understand these basic differences?
This can be comprehended by noting that these current quarks are not free but
are confined on account of the colour singlet character of the confining forces.
Thus we may state that these asymptotically free and very small-mass current
and relativistic quarks acquire higher constituent masses due to the fact that
they are confined. This conjecture turns out to be true in models where one
confines these small-mass current quarks within an appropriate potential or a
bag of appropriate shape and size. These potential and bag models are widely
discussed [42], [65].

271
272 Group Theory in Particle, Nuclear, and Hadron Physics

8.2 Confinement in a Spherically Static Bag


Consider a free quark with current mass defined by the Dirac equation,

Hψ = Eψ (8.1)
where

H = ᾱ · p̄ + βm (8.2)
ᾱ and β are standard Dirac operator.
Let us put the particle in a central potential U(r) which is bracketed with
the mass m as

α · p~ + βm + βU (r)]ψ(~x, t) = i
[~ ψ(~x, t) (8.3)
∂t
Now unlike in the non-relativistic case, the orbital angular momentum l is
not a good quantum number, even for a central potential.

————————————————–
Problem 8.1: Demonstrate that in the Dirac equation ~
 it is not l but the
total ~j = ~l + 12 Σ ~ = ~σ 0 .
~ which is connserved, where Σ
0 ~σ
————————————————–

It turns out that the conserved quantities of the above equations are J~2 ,
Jz , E and K where,
K = β(Σ ~ · ~l + 1).
Hψ = Eψ
J 2 ψ = j(j + 1)ψ
Jz ψ = µψ
Kψ = −κψ
ψ is a bispinor.
~ + 1)φ = −κφ , (~σ · L
(~σ · L ~ + 1)χ = κχ
where

!
  j3
φ gκ (r)Yjl (r̂)
ψjj3 (~r) = = j3 (8.4)
χ ifκ (r)Yjl 0 (r̂)

Note that l = j ± (1/2) and l0 = j ∓ (1/2) such that l + l0 = 2j.


From the above and noting that eigenstates of
< ~l · S
~ >= 1 < ~l · ~σ >= 1 (j(j + 1) − l(l + 1) − 3 ).
2 2 4
Thus K commutes with the Hamiltonian, the total angular momentum and
the spin. Hence the eigenvalues of (-κ) are given as
Bag Models 273
TABLE 8.1: Eigenvalues in a central
potential
κ jP l l0 State (n=1) (n=2)
-1 1/2+ 0 1 s1/2 2.04 5.40
1 1/2− 1 0 p1/2 3.81 7.00
-2 3/2− 1 2 p3/2 3.20 6.75
2 3/2+ 2 1 d3/2 5.12 8.41
-3 5/2+ 2 3 d5/2 4.33 8.06
3 5/2− 3 2 f5/2 6.37 9.75

κ = −(j + 12 ) = −l − 1 f or l = j − 21
= j + 21 = l f or l = j + 12
Consequently for j = 21 , κ = ±1. The lowest energy solution is given
by κ = −1 and l = 0, l0 = 1. We call it the 1s 12 state. n = 1 stands for
n − 1 = 0 nodes. The ‘s’ connotes that the upper component has zero angular
momentum (l=0) and j = 12 . For nodal excited states 2s 12 , 3s 21 , etc. with n=
2,3 . . . . Meanwhile, the same quantum numbers κ = −1, l = 0, l0 = 1 hold.
In Table 8.1 we provide the eigenvalues in a central potential, which are
in units of 1/R where R is the radius of the size of the central potential.
The lowest eigenvalues are:
2.04
E 1 + ,0 = (8.5)
R 2

For R ∼ 1f m. it is ∼ 400M eV , the mesonic mass is ∼ 800M eV and


the baryonic mass is ∼ 1200M eV . This is the same order of magnitude as
mρ ∼ 770M eV, mN = 940M eV and m∆ = 1232M eV . Note that a spin-orbit
force between same-l and different-j states exists.
Let us work out the state ψjj3 (r) in Equation 8.4 for the lowest states
explicitly. Now
       
j l 1/2 j j −1/2 1 l 1/2 j j3+1/2 0
Yj 3 = Yl 3 (r̂) + Yl (r̂) (8.6)
l j3 − 1/2 1/2 j3 0 j3 + 1/2 −1/2 j3 1

where the square brackets are the Clebsch-Gordan coefficients, well-tabulated


in the literature.
The lowest state κ = −1, l = 0, l0 = −1 is 1s1/2 with no nodes,

" j3
#
g−1 (~r)Yj=1/2,l=0 (r̂)
ψj=1/2,j3 = j3 (8.7)
if−1 (~r)Yj=1/2,l 0 =1 (r̂)

where
1 1 1 1
       
j3 l j −1 1 l j +1 0
Yj= = 1
2
1
2 Yl 3 2 (r̂) + 1
2 2 Yl 3 2 (r̂)
1
2 ,l j3 − 2 2 j3 0 j3 + 2 − 21 j3 1
(8.8)
274 Group Theory in Particle, Nuclear, and Hadron Physics

     
j3 =1/2 0 1/2 1/2 0 1 0 1
Yj=1/2,l=0 = Y (r̂) = Y0 (r̂) (8.9)
0 1/2 1/2 l=0 0 0

       
j =1/2
3 1 1/2 1/2 0 1 1 1/2 1/2 1 0
Yj=1/2,l=1 = Y1 (r̂) + Y1 (r̂) (8.10)
0 1/2 1/2 0 1 −1/2 1/2 1

r   r  
1 0 1 2 1 1
=− Y1 (r̂) + Y1 (r̂) (8.11)
3 0 3 0

Meanwhile, we find it eminently useful to exploit the following identity:

   
j1 j2 j3 j1 +j2 +j3 −2m3 j1 j2 j3
= (−1) (8.12)
−m1 −m2 −m3 m1 m2 m3

to obtain

     
j =−1/2 0 1/2 1/2 1 0
3
Yj=1/2,l=0 = Y10 (r̂) = Y00 (r̂) (8.13)
0 −1/2 −1/2 0 1

       
j3 =−1/2 1 1/2 1/2 −1 1 1 1/2 1/2 0 0
Yj=1/2,l=1 = Y (r̂) + Y (r̂)
−1 1/2 −1/2 1 0 0 −1/2 −1/2 1 1
r   r  
2 −1 1 1 0 0
=− Y (r) + Y (r) (8.14)
3 1 0 3 1 1
Thus

   
0 1
g−1 (~
r )Y0
  0q
 
ψj=1/2,j3 =1/2 =   (8.15)
1 0 
q
if−1 (~r)(− 13 Y10 (r) + 23 Y11 (r̂)

)
0 1

and

   
0
g−1 (r)Y00
 1 q
 
ψj=1/2,j3 =−1/2 =   (8.16)
1 0
q
if−1 (~r)(− 23 Y1−1 (r̂) + 13 Y10 (r̂)
 
)
0 1

————————————————–
Bag Models 275

Problem
   8.2: Let us define
 P ↑ as a projection operator such that
1 1 0
P↑ = with P ↑ = 0, and similarly for P ↓ . Then derive the
0 0 1
states obtained by the action of P ↑ and P ↓ on all the j = 1/2 states.
————————————————–

Use
j3 j3
(~σ · r̂)Yjl = −Yjl 0 (8.17)
Then
 
gκ (~r) j3
ψj,j3 (~r) = Yjl (8.18)
−ifκ (~r)(~σ · r̂)
where
       
j l 1/2 j j −1/2 1 l 1/2 j j +1/2 0
Yj 3 = Yl 3 (r̂) + Yl 3 (r̂)
l j3 − 1/2 1/2 j3 0 j3 + 1/2 −1/2 j3 1
(8.19)

————————————————–
Problem 8.3: Prove the expression Equation 8.17 above.
————————————————–

8.3 MIT Bag Model


From Equation 8.5 we see that as R → ∞ the energy in the spherical bag
drops monotonically. Hence the bag may extend to infinity and thus there
is no confinement in the above spherical bag model. However as per QCD
there should be confinement in the bag model. In the MIT bag model this
confinement is ensured by introducing a bag pressure B from outside. However
this is performed in an ad-hoc manner.
The MIT bag model is constructed to model QCD. The local gauge in-
variance would demand that one includes gluons in the formalism. One may
include gluons, but to avoid double-counting these should be treated only in
the lowest order of the perturbation theory. However we shall ignore gluons in
our calculation and write the Lagrangian density for the MIT bag for a quark
state ψ as:
i
L = [ (ψ̄γ µ ∂µ ψ − (∂µ ψ̄)γ µ ψ) − B]θV − (1/2)ψ̄ψδs (8.20)
2
with the step function
θV (r) = 1, inside the bag,
276 Group Theory in Particle, Nuclear, and Hadron Physics

θV (r) = 0, outside the bag, and with δs as the surface-delta function and B
is the bag constant.
Let us choose η µ as the outward normal for a static spherical bag. η µ =
(0, r̂), and minimizing the action associated with the above Lagrangian
∂L
= ((i/2)γ µ ∂µ ψ)θV − (1/2)ψδs (8.21)
∂ ψ̄
∂L
∂µ = (−(i/2)γ µ ∂µ ψ)θV − (i/2)γ µ ψδs (8.22)
∂(∂µ ψ)
Substituting these in the Lagrangian equation of motion
∂L ∂L
− ∂µ =0 (8.23)
∂ ψ̄ ∂(∂µ ψ)
and putting θV and δs term to zero, we get:
iγ µ ∂µ ψ = 0 ; inside the bag
iγ µ ∂µ ψ = ψ ; on the bag surface
The second is a linear boundary condition on the MIT bag surface, while
the first one is just the Dirac equation inside the bag for the massless quark.
Consider now the energy-momentum tensor
∂L ∂L
T µν = ( ∂ ν ψ + ∂ ν ψ̄ ) − g µν L
∂(∂µ ψ) ∂(∂µ ψ)
= (i/2)(ψ̄γ µ ∂ ν ψ − ∂ ν ψ̄γ µ ψ)θν − g µν L (8.24)

∂µ T µν = (i/2)∂µ (ψ̄γ µ ∂ ν ψ − ∂ ν ψ̄γ µ ψ)θV − ∂µ g µν L (8.25)


For the last term
i
−∂µ g µν L = −∂ ν L = (−∂ ν ψ̄γ µ ∂µ ψ + ∂µ ψ̄γ µ ∂ ν ψ)θV
2
1
+B∂ ν θV (x) + ∂ ν (ψ̄ψδs ) (8.26)
2
Next the first term in bracket

(i/2)(−∂ ν ψ̄γ µ ∂µ ψ + ∂µ ψ̄γ µ ∂ ν ψ)θν + (i/2)(ψ̄γ µ ∂ ν ψ − ∂ ν ψ̄γ µ ψ))∂µ θV (8.27)

use ∂ µ θV = ηµ δs .
Putting together

∂µ T µν = (i/2)(−2∂ ν ψ̄γ µ ∂µ ψ + 2∂µ ψ̄γ µ ∂ ν ψ)θV + Bδs η ν


+(1/2)∂ ν (ψ̄ψδs ) + (i/2)(ψ̄γ µ ∂ ν ψ − ∂ ν ψ̄γ µ ψ)ηµ δs
Note the first term above
Bag Models 277

−2∂ ν ψ̄γ µ ∂µ ψ = −2g µν ∂µ ψ̄γ µ gµν ∂ ν ψ = −2∂µ ψ̄γ µ ∂ ν ψ(g µν gµν )

= −2∂µ ψ̄γ µ ∂ ν ψ (8.28)


And thus this cancels the other term in the first bracket. Conservation of
energy-momentum implies ∂µ T µν = 0 and thus the following hold separately:

Bηs η µ + (i/2)(ψ̄γ µ ∂ ν ψ − ∂ ν ψ̄γ µ ψ)ηµ δs = 0 (8.29)


and ∂µ (ψ̄ψδs ) = 0.
Thus
Bηs η ν = (1/2)(∂ ν ψ̄iγ µ ηµ ψ − iψ̄γ µ η µ ∂ ν ψ)δs
= (1/2)((∂ ν ψ̄)ψ + ψ̄(∂ ν ψ))δs = (1/2)∂ ν (ψ̄ψ)δs (8.30)
so Bη ν = (1/2) ∂x∂ ν (ψ̄ψ) ; (on the surface).
It is the introduction of B, which we may now call the bag pressure, that
ensures energy-momentum conservation. As η µ ηµ = −1 (η µ is a space-like
four-vector), one obtains on the surface the non-linear boundary condition of
the MIT bag:
B = −(1/2)ην ∂ ν (ψ̄ψ)
For a spherical state boundary condition η ν = (0, r̂).
It can be shown that
B = −(1/2) ∂x∂ ν (ψ̄ψ)
which is bag pressure inwards from outside and which balances the outward
pressure of the quarks from inside on the boundary.

————————————————–
Unsolved Problem 8.1: Demonstrate that the outward pressure of
quarks from inside a bag is equal to the bag pressure B.
————————————————-
Problem 8.4: Given T µν above, show that ηµ T µν = 0 on the surface of
the bag. This ensures that no energy escapes the bag boundary.
————————————————–

Now given ηµ T µν = 0 and noting that the T µν defined here is the most
general Dirac definition of the energy momentum tensor, we can therefore
define the energy-momentum tensor of the bag as

TBµν (x) = [T µν (x) − Bg µν ]θV (x) (8.31)


Thus the energy-momentum four-vector is
Z Z
Aν = d3 xTB0ν (x) = d3 x(T 0ν (x) − Bg 0ν ) (8.32)
V
278 Group Theory in Particle, Nuclear, and Hadron Physics

where V is the volume of the bag over which the integration is performed.
The total bag momentum and energy are, respectively,
Z
Pi = d3 xT 0i (x) (8.33)
V
Z
E = P0 = d3 xT 00 (x) − BV (8.34)
V
where V is the bag volume.
Note that the bag pressure B contributes to the energy but does not con-
tribute to the momentum. Furthermore, observe that the pressure B arises
from the structure of the vacuum of the underlying theory of the strong in-
teraction, i.e., QCD.
Now we saw earlier that in the ground state of a quark bound by a spher-
ically symmetric potential its energy range is 2.04/R. So in the above we put
N number of quarks. Then its energy becomes

N × 2.04 4πR3 B
E= + (8.35)
R 3
Let us minimize it as a function of R,
∂E N × 2.04
=0=− + 4πR2 B (8.36)
∂R R2

(N × 2.04)1/4
R= (8.37)
(4πB)1/4
Substituting it back into Equation 8.35 above we obtain the energy or
mass of N confined quarks,
4
MN C 2 = (4πB)1/4 (N × 2.04)3/4 (8.38)
3
Alternatively, we obtain R = 43 2.04N
E .
From the above, as the meson has 2 quarks and the baryon has 3,
M eson M2 2
= = ( )3/4 (8.39)
Baryon M3 3
And the size of a nucleon for N = 3 we get R = 1.6f m.
Thus size generates mass of a proton from the mass of quarks. This mass
arises due to the fact that the quarks have kinetic energy and are confined.
So this process has generated a constituent mass of quarks of the order of
340 M eV from its current quark mass of zero.

Bag Model Wave Function of Nucleons


We saw in the above that the state of a quark confined in a static spherical
bag is given as:
Bag Models 279

!
  j3
φ gκ (r)Yjl (r̂)
ψjj3 (~r) = = j3
χ ifκ (r)Yjl 0 (r̂)

And then we obtained the expressions for the state j = 1/2, j2 = 1/2 and
j = 1/2, j3 = −1/2.
As this is for a bound quark, it is a flavour state in terms of the quark
model as we discussed earlier. In the non-relativistic case as SU (6) = SU (3) ×
SU (2), we obtain SU (6)sf symmetry states by taking the product of the
appropriate SU (3)f state and the SU (2)s state for bound configurations of
(q q̄) and (qqq). We saw that the SU (6)sf fully symmetric state describes the
hadronic configuration well. We build a fully symmetric state as:
1
Ψ = √ (ψρ φρ + ψx φx ) (8.40)
2
where ψρ and ψx are mixed anti-symmetric and mixed symmetric states in
the flavour space and φρ and φλ the corresponding states in the spin space.
Note the full space was SU (3)f × SU (2)s × SO(3)l where SO(3)l describes
the orbital space. One takes appropriate, say l = 0 state for the ground state.
But this luxury of separation of spin and orbital states does not exist in the
relativistic Dirac equation. Here the space and spin parts mix up. Hence one
may assume that the proper group structure for the bag models SU (3)f ×
(SU (2)s × SO(3)l ) and the states ψj=1/2,j3 =1/2 and ψj=1/2,j3 =−1/2 may be
attributed as “spin” state of the group (SU (2)s × SO(3)l ).
And thus we replace spin up |j = 1/2, j3 = 1/2 > and spin down |j =
1/2, j3 = −1/2 > states of non-relativistic quark model of the group SU (2)
with ψj=1/2,j3 =1/2 and ψj=1/2,j3 =1/2 states of the group (SU (2)s × SO(3)l )
which we assume is inseparable,
Remember in our notation φρp = √12 (ud−du)u and φλp = − √16 (udu+duu−
2uud) and hence
1
ψ ρ1 1 = √ (ψ 21 12 ψ 12 − 12 − ψ 21 − 12 ψ 12 12 )ψ 12 12 (8.41)
2 2 2

1
ψ λ1 1 = − √ (ψ 21 12 ψ 12 − 12 ψ 12 21 + ψ 12 − 12 ψ 21 12 ψ 21 12 − 2ψ 21 12 ψ 21 12 ψ 21 − 12 ) (8.42)
2 2 6

1
ψ ρ1 − 1 = √ (ψ 21 12 ψ 12 − 12 − ψ 21 − 12 ψ 12 12 )ψ 12 − 12 (8.43)
2 2 2

1
ψ λ1 − 1 = √ (ψ 12 12 ψ 12 − 21 ψ 12 − 21 + ψ 12 − 21 ψ 12 12 ψ 12 − 21 − 2ψ 12 − 21 ψ 12 − 12 ψ 12 12 ) (8.44)
2 2 6
Now our group in the MIT bag model is SU (3)f × (SU (2)s × SO(3)l ) and
thus we consider a general operator of the form
280 Group Theory in Particle, Nuclear, and Hadron Physics

3
X
O= Ii Oi (~ri , ~σi ) (8.45)
i=1

where Ii is an operator on flavour space SU (3)f and Oi (ri , σi ) is an operator


with space SU (2)f ⊗ SO(3)l space on state ψjj3 of the bag as these have both
orbital and spin degrees of freedom mixed. It acts on the nucleon state Ψ
defined above. Now as the state Ψ is symmetric,

3
X
< Ψ|O|Ψ >=< Ψ| Ii Oi (~ri , ~σi )|Ψ >= 3 < Ψ|I3 O3 (~ri , ~σi )|Ψ > (8.46)
i=1

where 3 subscript denotes the operator acting on quark number three in a


nucleon,

3
< Ψ|O|Ψ >= < φρ ψρ + φλ ψλ |I3 O3 (~ri , ~σi )|φρ ψρ + φλ ψλ > (8.47)
2
If I3 is an operator which does not change the flavour of quark, it acts on
these as < φρ |I3 |φλ >= 0. Then

3 3
< ψ|O|ψ >= < φρ |I3 |φρ >< ψρ |O3 |ψρ > + < φλ |I3 |φλ >< ψλ |O3 |ψλ >
2 2
(8.48)
Let us take the N with spin up N ↑ for the Ψ state
3 ρ ρ
< ΨN↑ |O|ΨN↑ >= < φN N
ρ |I3 |φρ >< ψ 1 1 |O3 |ψ 1 1 >
2 2 2 2 2

3 ρ ρ
+ < φN N
λ |I3 |φλ >< ψ 1 1 |O3 |ψ 1 1 > (8.49)
2 2 2 2 2

Now
ρ ρ
< ψ1/2,1/2 |O3 |ψ1/2,1/2 >

1 3 3
= (< ψ1/2,1/2 ψ1/2,1/2 |ψ1/2,1/2 ψ1/2,−1/2 >< ψ1/2,1/2 |O3 |ψ1/2,1/2 >
2

3 3
+ < ψ1/2,−1/2 ψ1/2,1/2 |ψ1/2,−1/2 ψ1/2,1/2 >< ψ1/2,1/2 |O3 |ψ1/2,1/2 >) (8.50)

where we have used label ‘3’ for the state of the third quark. Using orthogo-
3 3
nality of the states this is equal to < ψ1/2,1/2 |O3 |ψ1/2,1/2 >.

λ λ 1 3 3 3 3
< ψ1/2,1/2 |O3 |ψ1/2,1/2 >= (< ψ1/2,1/2 |O3 |ψ1/2,1/2 > + < ψ1/2,1/2 |O3 |ψ1/2,1/2 >
6
Bag Models 281
3 3
+4 < ψ1/2,−1/2 |O3 |ψ1/2,−1/2 >) (8.51)
Putting these together, we obtain:

< ΨN↑ |Σ3i=1 Ii Oi (ri , σi )|ΨN↑ >=< φN N 3 3


λ |I3 |φλ >< ψ1/2,−1/2 |O3 |ψ1/2,−1/2 >

3 1
+ (< φN N N N 3 3
ρ |I3 |φρ > + < φλ |I3 |φλ >) < ψ1/2,1/2 |O3 |ψ1/2,1/2 > (8.52)
2 3
This is the main expression which we shall use to obtain different observ-
ables in the MIT bag model.

Normalization of the Wave Function


We see that
 
gκ (r)
ψjj3 (r) = Yjjl3 (8.53)
−if−κ~σ · r̂
It turns out that this state as a bound state is given as [65] for the ground
state w = 2.04, E = 2.04/R
 
N j0 (wr/R)
ψκ=−1 = χµκ=−1 (8.54)
(4π)1/2 i~σ · r̂j1 (wr/R)
The spin angle function as χµk=−1 are the upper components in an s-state.
This is for r < R and zero for r > R.
Now we want to determine the normalization constant N :

Z Z R  
∗ 2 j0 (wr/R)
ψ ψdr = N (j0 (wr/R) − i~σ · r̂j1 (wr/R)) < χµ |χµ >
0 i~σ · r̂j1 (wr/R)
(8.55)
2
< χµ |χµ >= 1 and as (~σ · r̂) = 1
Z Z R
∗ 2
ψ ψdr = N (|j0 (wr/R)|2 + |j1 (wr/R)|2 )r2 dr (8.56)
0
R
With wr/R = x then dx = w dr.
Take the first term:

R w w
R3
Z Z Z
R
(j0 (wr/R))2 r2 dr = J02 (x)(Rx/w)2 dr = 3 J02 (x)x2 dx (8.57)
0 0 w w 0

Then one obtains,


282 Group Theory in Particle, Nuclear, and Hadron Physics

R3 3 2 w R3 2
= (x /2[j0 (x) + η 0 (x)j 1 (x))] 0 = (j (w) + η0 (w)j1 (w)) (8.58)
w3 2 0
Next
R w
R3
Z Z
(j1 (wr/R))2 r2 dr = j12 (x)x2 dx
0 w3 0

R3 3 2 w R3 2
= [x /2(j1 (x) − j 0 (x)j 2 (x))] 0 = (j (w) − j0 (w)j2 (w)) (8.59)
w3 2 1
The limit m → ∞ implies that the quarks behave as if they were free
inside R with an effective mass of zero and exhibit infinite mass outside. In
other words, they are confined. In that case matching the derivative inside
and outside leads to the condition [42],

J0 (w) = j1 (w) (8.60)

Z R
(|j0 (wr/R)|2 + |j1 (wr/R)|2 )r2 dr
0

R3 2
= [(j1 (w) + η0 (w)j1 (w)) + (j12 (x) − j0 (x)j2 (x))]
2
R3
= j0 (w)[2j0 (w) + η0 (w) − j2 (w)] (8.61)
2

R3 sin w sin w cos w 3 1 cos w


= [2 − − ( 3 − ) sin w + 3 2 ] (8.62)
2 w w w w w w
2R3
Thus N 2 = w3 sin2 w(w − 1) = 1. So,

w3
N2 = (8.63)
2R3 (w − 1) sin w
However a note of warning. The above normalization condition is for the
special condition Equation 8.63. However it is not essential. Several groups do
not impose this condition and then a different normalization term arises. This
term is more popular in the current literature. However in the older literature
yet another normalization is imposed which we use here now.
Let us continue with Equation 8.61
Z R
(|j0 (wr/R)|2 + |j1 (wr/R)|2 )r2 dr
0

R3 2
= [(j1 (w) + η0 (w)j1 (w)) + (j12 (x) − j0 (x)j2 (x))]
2
Bag Models 283

R3
= [j0 (w)(j0 (w) − j2 (w)) + j1 (w)(η0 (w) + j1 (w))]
2

R3 sin w sin w 3 1 3 cos w


= [ ( − ( 3 − ) sin w + )
2 w w w w w2

sin w cos w cos w sin w cos w R3 sin2w


+( − )(− + − )] = (1 − )=1 (8.64)
w2 w w w2 w w2 w2
Thus

w4
N2 = (8.65)
R3 (w2 − sin2w )
All is well as long as we use these normalization terms consistently.

The Axial Vector Coupling Constant


The axial vector coupling constant is defined as [42]

< p↑ | i τi+ Siz |n↑ >


P
gA
= (8.66)
gV < p↑ |Σi τi+ |n↑ >
We saw earlier that in the non-relativistic quark model the above evaluates
to 5/3. Thus what we saw was considered a failure of the non-relativistic quark
model, as the empirical value is ∼ 1.1. In the previous chapter we solved the
problem by assuming that the nucleon is deformed in the ground state. Here
we see if the relativistic corrections may solve the same problem.
Now a quark-spin matrix element < q|σz |q > is as in the non-relativistic
quark model. However in the relativistic case, the upper l = 0 component and
the lower l = 1 component and thus < q|σz |q > is no longer unity as the spin
may come from the lower component. Hence one may have
gA 5
= < σz > (8.67)
gV 3
where < σz > is the expectation value of the spin-z projection of the rel-
ativistic quarks, Let us define the total angular momentum of a quark as
jz = 21 σz + lz so < σz >= 2 < jz > −2 < lz >. Using the state in Equation
8.54
A
< jz >= (8.68)
B
where
Z R
j =1/2 j =1/2
A = 4π |j0 (wr/R)|2 r2 dr < Yj=1/2,l
3 3
|jz |Yj=1/2,l >
0
284 Group Theory in Particle, Nuclear, and Hadron Physics

Z R
2 j =1/2 j =1/2
+4π |j1 (wr/R)|2 r2 (~σ · r̂) dr < Yj=1/2,l
3 3
|j3 |Yj=1/2,l > (8.69)
0

and
Z R Z R
B = 4π( |j0 (wr/R)|2 r2 dr + |j1 (wr/R)|2 r2 dr) (8.70)
0 0
Note that

3 j =1/2 3 j =1/2 1 1
< Yj=1/2,l |jz |Yj=1/2,l >= , < jz >=
2 2
Next
   
j =1/2 Y00 j3 =1/2 −Y
3
Yj=1/2,l=0 = and Yj=1/2,l=1 = √ 00
0 2Y11
Thus
3 j =1/2 3 j =1/2
< Yj=1/2,l=0 |lz |Yj=1/2,l=0 >= 0
and

3 j =1/2 3 j =1/2 1 2
< Yj=1/2,l=1 |lz |Yj=1/2,l=1 >= < Y10 |lz |Y10 > + < Y11 |lz |Y11 >
3 3

2 2
=0+ < Y11 |Y11 >=
3 3
Thus
RR
(2/3) |j1 (wr/R)|2 r2 dr
0
< lz >= R R RR (8.71)
0
|j0 (wr/R)|2 r2 dr + 0 |j1 (wr/R)|2 r2 dr
R3 2 sin2 w R3
The denominator is 2 w2 (1 − w2 ) = w4 (w
2
− sin2 w)
Now
R
R3 2
Z
|j1 (wr/R)|2 r2 dr = (j (w) − j0 (w)j2 (w))
0 2 1
3
R sin w cos w 2 sin w 3 3
= [( 2 − ) − (( 2 − 1/w) sin w − 2 cos w)]
2 w w w w w
R3 1 2 sin2 w sin w cos w
= . 2 [1 − + ] (8.72)
2 w w2 w
sin w(w cos w−sin w)
Thus < lz >= (1/3)[1 + w2 −sin2 w
]
Using j0 (w) = j1 (w)
Bag Models 285

Then sinww = sin w


w2 −
cos w
w
2w−3
lz = (1/6) w−1
Thus,

gA 2w − 3
= (5/3) < σz >= (5/3)(1 − 2 < lz >) = (5/3)(1 − ) (8.73)
gv 3(w − 1)

For w = 2.04, get ggAv = 1.1, which puts the experimental results well.
Thus success in fitting ggAv well has conventionally been taken as a big
success of the MIT bag model. However as we shall see below, the MIT bag
model fails to describe the spin dependent structure function of nucleons. We
have already seen above that the deformed nucleon is successful in this case
too.

The Magnetic Moment in the MIT Bag Model


Define the magnetic moment operator,
Z
1
µ= ~r × ~jem dr (8.74)
2
Z
1 X †
µ= d~r ~r × [ qi (~r)~
αi Qi qi (~r)] (8.75)
2 i
 
0 σ
with α
~=
σ 0
and
 
N j0 (wr/R)
q(~r) = ψn,κ=−1 = χµ1/2 (8.76)
(4π)1/2 i~σ · r̂j1 (wr/R)
w3
here N 2 = 2R3 (w−1) sin w as we saw earlier.
So,
Z R
N2 X
µ= 2π Qi r2 dr(j0 (wr/R) − iσ · r̂j1 (wr/R)·
4π i 0
  
0 ~σ j0 (wr/R)
~r ×
~σ 0 i~σ · r̂j1 (wr/R)
Z R
N2 X
µ= Qi r2 dr[i~r ×~σ~σ · r̂j0 (wr/R)j1 (wr/R)−i~σ · r̂~r ×~σ j0 (wr/R)j1 (wr/R))]
2 0
Z R
N2 X
µ= Qi r2 drij0 (wr/R)j1 (wr/R)[(~r ×~σ )(~σ ·r̂)−(~σ ·r̂)(~r ×~σ )] (8.77)
2 0
Let us simplify
286 Group Theory in Particle, Nuclear, and Hadron Physics

(~r × ~σ )(~σ · r̂) − (~σ · r̂)(~r × ~σ )


The first term is

~σ · ~r
(~r × ~σ )(~σ · r̂) = (~r × ~σ )
r
This is equal to

(~r × ~σ
(~σ · r̂)
r

= (1/r)[i(yσz − zσy ) − j(xσz − zσx ) + k(xσy − yσx )](xσx + yσy + zσz )

The first term in above becomes

(yσz − zσy )(xσx + yσy + zσz ) = i[x(xσx + yσy + zσz ) − r2 σx ] = i[x~r · ~σ − γ 2 σx ]

Similar expressions by symmetry for the other terms yield

(~r × ~σ )~σ · ~r = (i/r)[ix~r · ~σ + jy~r · ~σ + kz~r · ~σ − r2 iσx − r2 iσy − r2 iσz ]

= (i/r)[~r(~r · ~σ ) − r2~σ ] = i[r̂(~r · ~σ ) − r~σ ]


The other terms now is
1
(~σ · r̂)(~r × ~σ ) = (~σ · ~r)(~r × ~σ (8.78)
r

= (1/r)(xσx + yσy + zσz )[i(yσz − zσy ) − j(xσz − zσx ) + k(xσy − yσx )] (8.79)

The first term becomes

(xσx + yσy + zσz )(yσz − zσy ) = i[r2 σx − x~r · ~σ ]


Thus by symmetry substitution for the other two,

(~σ · r̂)(~r × ~σ )

~ = i[r~σ − r̂(~r · ~σ )]
= (i/r)[r2~σ − ~r(~r · σ)]
Thus

(~r × ~σ )(~σ · r̂) − (~σ · r̂)(~r × ~σ ) = 2i[r̂(~r · ~σ ) − r~σ ] (8.80)


Hence
Bag Models 287

Z R
N2 X
µ= Qi r2 dr [i(2i)]j0 (wr/R)j1 (wr/R)(r̂(~r · ~σ ) − r~σ ) (8.81)
2 i 0

Note that ~r~r here gives (1/3)r2 in the scalar part. In the
R
d~r integration
only this will survive. And so it separates out the r~σ term.

R
N2 X
Z
µ = (2/3) 2 Qi < ~σi > r3 j0 (wr/R)j1 (wr/R)dr (8.82)
2 i 0

When for the ground state < σi >= 1 and next


R w
R4
Z Z
3
r j0 (wr/R)j1 (wr/R)dr = 4 x3 j0 (x)j1 (x)dx (8.83)
0 w 0

where
R wr/R = xR R dv R
Use vudx = v( udx) − dx ( udx)dx
Z Z
x3 j0 (x)j1 (x)dx = (xj1 (x))x2 j0 (x)dx
Z Z Z
2 d
= xj1 (x) x j0 (x)dx − ( xj1 (x)) x2 j0 (x)dx)dx
dx
Z
d
= xj1 (x)(x2 j1 (x)) − [j1 (x) + x j1 (x)]x2 j1 (x)dx
dx
Z
= x3 j12 (x) − [j1 (x) + x(j0 (x) − 2/xj1 (x))]x2 j1 (x)dx
Z Z
= x3 j12 (x) − 3
x j0 (x)j1 (x)dx + x2 j1 (x)dx

We see
x3 2
Z
x2 j1 (x)dx = [j (x) − j0 (x)j2 (x)]
2 1
w3 2
Z
3
x3 j0 (x)j1 (x)dx = [ x3 j1 (x)−1/4x3 j0 (x)j1 (x)]w
0 = [3j1 (w)−j0 (w)j2 (w)]
4 4
Using the boundary condition, j0 (w) = j1 (w).
The right-hand side in above is
w3 w3 sin w sin w 4w−3
4 j0 (w)[3j0 (w) − j2 (w)] = 4 w ( w ( w ))
Z w
1
x3 j0 (x)j1 (x)dx = sin2 w(4w − 3) (8.84)
0 4
And hence the magnetic moments of massless quarks in the ground state
is
288 Group Theory in Particle, Nuclear, and Hadron Physics

R 4w − 3
µ= eq (8.85)
12 w(w − 1)
e
For x = 2.04, µ = 0.203eq R = 0.203(eq /e)eR = 0.203(eq /e)(2mp R) 2m p
e
In units of nuclear magneton 2m p
(where m p is mass of proton) and R
expressed in fermis µ = 1.93(eq /e)RP
For proton adding quark charge eq = e and so the magnetic moment of
proton (in units of nuclear magneton) Mp = 1.93R.
For R = 1f m, Mp = 1.93, compare this to the experimental value of 2.79
nm. So in the MIT bag model it is not possible to fix nuclear radius and
magnetic moment simultaneously.

————————————————–
Problem 8.5: Determine the matrix element of the axial current as

Z Z

< p|qi† γ 0~γ γ 5 qi (r)|n >
X
3 ~
d r < p|A(r)|n > |k=0 = d3 r
i
2

and see the connection with ggVA , the axial vector coupling constant obtained
by using the expression for the total angular momentum.
————————————————–

Spin of the Nucleon in the MIT Bag Model


Let us now derive the expression for the integrated spin dependent struc-
ture function of proton and neutron as defined earlier [66]. We shall use the
matrix element of the operator,
3
X
O= Ii Oi (~ri , ~σi ) (8.86)
i=1

where Ii is an operator in SU (3)f space and Oi in SU (2)s × SO(3)l space.


For nucleon with spin up the matrix element is
3
X
< ψN ↑ | Ii Oi (~ri , ~σi )|ψN ↑ >
i=1

=< φN N
λ |I3 |φλ >< ψ1/2,−1/2 (3)|O3 |ψ1/2,−1/2 (3) >

3 1
+ (< φN N N N
ρ |I3 |φρ > + < φλ |I3 |φλ >) < ψ1/2,1/2 (3)|O3 |ψ1/2,1/2 (3) >
2 3
(8.87)
where ψ1/2,1/2 and ψ1/2,−1/2 states from eqns. (41)-(44).
As we shall be requiring values of u↑ ,u↓ , d↑ and d↓ , etc. in our calculations,
Bag Models 289

let us define I3 = Pu (3) or Pd (3), where Pu (3) picks up up-quark states and
Pd (3) picks up down-quark states sitting in position 3. Also Pu (3) in a d-quark
gives zero and Pd (3) gives zero acting on a u-quark in position 3.
We also have from the previous results,

< φpρ |Pu (3)|φpρ >= 1, < φpρ |Pd (3)|φpρ >= 0,
< φpλ |Pu (3)|φpλ >= 1/3, < φpλ |Pd (3)|φpλ >= 2/3 (8.88)
Let us check for O(3) operator P ↑ (3) and P ↓ (3) which picks up spin-up
state and spin down states, respectively, at position 3 and zero for the others.
This operator has already been used in a problem set earlier.
We are assuming that there are no s-quarks and gluons in our model here.
We need for example

u↑ =< P ↑ |Pu (3)P ↑ (3)|P ↑ > (8.89)


In expression Equation 8.87 above

< P ↑ |Pu (3)P ↑ (3)|P ↑ >=< φpλ |Pu (3)|φpλ >< ψ1/2,−1/2 (3)|P ↑ (3)|ψ1/2,−1/2 (3) >

3 1
+ (< φpρ |Pu (3)|φpρ > + < φpλ |Pu (3)|φpλ >) < ψ1/2,1/2 (3)|P ↑ (3)|ψ1/2,1/2 (3) >
2 3
(8.90)
Let us calculate < ψ1/2,1/2 (3)|P ↑ (3)|ψ1/2,1/2 (3) > which is a one quark
matrix element.
We showed for example that given,

   
0 1
g−1 (~
r )Y0
 0
 
ψj=1/2,j3 =1/2 =   (8.91)
 p 1 p 0 
if−1 (~r)(− 1/3Y10 (r̂) + 2/3Y11 (r̂ )
0 1

And so
   
1
g−1 (~r)Y00
0  
P ↑ ψ1/2,1/2

=  (8.92)
 p 0 1 
if−1 (~r)(− 1/3Y1 (r̂) )
0
Hence

< ψ1/2,1/2 (3)|P ↑ (3)|ψ1/2,1/2 (3) >


290 Group Theory in Particle, Nuclear, and Hadron Physics

   †    
1 1
g−1 (~r )Y00 r )Y00
g−1 (~
2 0   0 
= N (         )
1 0   1 
if−1 (r)(− 1/3Y10 (r̂) + 2/3Y11 (r̂ r )(− 1/3Y10 (r̂)
 p p p
) if−1 (~ )
0 1 0

Likewise,
Z R Z R
2 2 2
=N ( r (g− (r)) dr + (1/3) r2 (f−1 (r))2 dr)
0 0
Z R Z R
2 2
=N ( r j0 (wr/R)dr + (1/3) r2 j1 (wr/R)dr) (8.93)
0 0
The last expression is for the MIT bag model.
Now we have

w3
N2 = (8.94)
2R3 (w − 1) sin2 w
And
R
R3 sin2 w(2w − 1)
Z
r2 j02 (wr/R)dr = (8.95)
0 2 w3
R
R3 sin2 w(2w − 3)
Z
r2 j12 (wr/R)dr = (8.96)
0 2 w3
Therefore
4w − 3
< ψ1/2,1/2 (3)|P ↑ (3)|ψ1/2,1/2 (3) >= (8.97)
6(w − 1)
Similarly,

< ψ1/2,1/2 (3)|P (3)|ψ1/2,1/2 (3) >
   †
1
r )Y00
 
g−1 (~ 0
2 0   
= N ( 0 )
r )( 2/3Y11 (r̂)
      p
1 0  if−1 (~ )
if−1 (r)(− 1/3Y10 (r̂) + 2/3Y11 (r̂)
 p p
) 1
0 1
Z R Z R 2w − 3
2 2 2 2 2 2
=N r (2/3)j−1 (r)dr = N r (2/3)j1 (wr/R)dr = (8.98)
0 0 6(w − 1)

One finds
2w − 3
< ψ1/2,−1/2 (3)|P ↑ (3)|ψ1/2,−1/2 (3) >= (8.99)
6(w − 1)
4w − 3
< ψ1/2,−1/2 (3)|P ↓ (3)|ψ1/2,−1/2 (3) >= (8.100)
6(w − 1)

————————————————–
Bag Models 291

Unsolved Problem 8.1: Compare, by explicit calculations, the above


expressions for the two matrix elements.
————————————————–

Putting it all together in Equation 8.87, we obtain

2w − 3 3 4w − 3
u↑ =< p↑ |Pu (3)P ↑ (3)|p↑ >= 1/3( ) + (1 + 1/9)( )
6(w − 1) 2 6(w − 1)
11w − 9
= (8.101)
9(w − 1)
and
u↓ =< p↑ |Pu (3)P ↓ (3)|p↑ >=< φpλ |Pu (3)|φpλ >< ψ1/2,−1/2 (3)|P ↓ (3)|ψ1/2,−1/2 (3) >

3 1
+ (< φpρ |Pu (3)|φpρ > + < φλp |Pu (3)|φλp >) < ψ1/2,1/2 (3)|P ↓ (3)|ψ1/2,1/2 (3) >
2 3
7w − 9
= (8.102)
9(w − 1)
Check u = u↑ + u↓ = 2.
It shows that there are two up-quarks in the proton. Also we find
8w − 9
d↑ =< p↑ |pd (3)p↑ (3)|p↑ >= (8.103)
18(w − 1)
10w − 9
d↓ =< p↑ |pd (3)p↓ (3)|p↑ >= (8.104)
18(w − 1)
Note d = d↑ + d↓ = 1. Showing that there is only one down quark in the
proton.

————————————————–
Unsolved Problem 8.2: Confirm these two expressions for d↑ and d↓ by
performing explicit calculations analogous to those illustrated above.
————————————————–

Note that for w = 2.04,

u↑ = 1.436, u↓ = 0.564, d↑ = 0.391, d↓ = 0.609 (8.105)


Now the integrated spin-dependent structure function of proton is (see
next chapter for details)

Z 1 Z 1
1 4 1
g1p = g1p (x) = dx[ (u↑ (x) − u↓ (x)) + (d↑ (x) − d↓ (x))]
0 2 0 9 9
292 Group Theory in Particle, Nuclear, and Hadron Physics
1 4 ↑ 1
= [ (u − u↓ ) + (d↑ − d↓ )] (8.106)
2 9 9
5 w
g1p = ( ) (8.107)
18 3(w − 1)
For w = 2.04 we get g1p = 0.182
Note that the non-relativistic value of 5/18 for g1p is being relativistically
corrected by the same factor as for gA = 5/3, as was accomplished earlier. So
when ggAv = 1.01 correctly fixed the above g1p , the final value of 0.182 is far
larger than the EMC experimental value of 0.123 [56]. Thus if fixing of ggAv
as 1.01 in the MIT bag model was taken as success of the MIT bag model
confinement concept of the relativistic quarks, then clearly this lack of fitting
of g1p should be taken as a failure of the MIT bag model.
Next by u ↔ d exchange symmetry between the neutron and proton wave
functions, we obtain the neutron spin structure function as:
4 1
g1n = [ (d↑ − d↓ ) + (u↑ − u↓ )] = 0 (8.108)
9 9
in the MIT bag model. Now these g1p and g1n and gA /gV values in the MIT
bag model do satisfy the Bjorken sum rule,
1 gA
g1p − g1n = ( ) (8.109)
6 gV
However this prediction of MIT bag model for g1n = 0 is in gross conflict
with the experimental value of g1n [57], [67], [68] of -0.31. -0.037, and -0.058,
respectively. On the other hand, the deformed nucleon model, as we demon-
strated in the previous chapter, matches these experimental values pretty well.

8.4 Finite Mass Quarks in a Bag


For finite quarks of mass m confined in a bag, for the ground state the
wave function is
 q 
E+m
N (w)  E j0 (wr/R)
g(r) = √ q  (8.110)
4π E+m
i~σ · r̂j1 (wr/R)
E

w2 +m2 R2
where E = R
This massive quark system will permit us to study the two limiting cases:
that of the extreme relativistic quarks (m = 0) which we have already an-
alyzed in detail here, and next that of the non-relativistic quarks which we
investigated in the previous chapter.
Note the boundary condition in the bag is
Bag Models 293

E+m 2 E−m 2
( )j0 (w) − ( )j1 (w) = 0
m m
Whence,
r
E−m
j0 (w) = j1 (w) (8.111)
E+m
Next we obtain the normalization tern N(w)

R
E−m 2
Z Z
E+m 2
q ∗ qd~r = N 2 r2 [( )j0 (wr/R) + ( )j1 (wr/R)]dr
0 m m
Z R Z R
m
= N 2[ r2 (j02 (wr/R) + j12 (wr/R)) + r2 (j02 (wr/R) − j12 (wr/R))]
0 E 0
(8.112)
Now
R
R3 2
Z
r2 j02 (wr/R)dr = [j (w) + η0 (w)j1 (w)]
0 2 1
R3 E − m 2
= [( )j (w) + η0 (w)j1 (w)]
2 E+m 1
R3 j1 (m)
= [E(j1 (w) + η0 (w)) + m(−j1 (w) + η0 (w))]
2 E+m
Now
sin w
j1 (w) + η0 (w) = w2 − 2 cosww

−j1 (w) + η0 (w) = − sin


w2
w

For the q
boundary condition
sin w
w = E−m sin w cos w
E+m ( w2 − w )
Therefore q
cos w = sinww − E−mE+m sin w

R
r
R3 0 E−m
Z
2 1 2E
r j0 (wr/R)dr = j1 (w)[− + ] (8.113)
0 2 w E+m E−m
p
As w = R (E − m)(E + m)
Therefore
R
R3 2
Z
1 1
r2 j02 (wr/R)dr = j (w) (− + 2E) (8.114)
0 2 0 E−m R
Note for m → 0 , ER = w and then
294 Group Theory in Particle, Nuclear, and Hadron Physics

R
R3 2 2w − 1
Z
r2 j02 (wr/R)dr = j0 (w) (8.115)
0(m→0) 2 w
which checks out with our previous result. Next
R
R3 2
Z
r2 j12 (wr/R)dr = [j (w) − j0 (w)j2 (w)]
0 2 1
R3 2 1
= j0 (w) (2E − 3/R) (8.116)
2 E−m
For m → 0 , E = w/R and we obtain
R
R3 2 2w − 3
Z
r2 j12 (wr/R)dr = j (w) (8.117)
0 2 0 w
which checks out with our earlier result

Z R
1
r2 (j02 (wr/R) + j12 (wr/R))dr = 2R3 j02 (w) (E − 1/R) (8.118)
0 E−m
Z R
1
r2 (j02 (wr/R) − j12 (wr/R))dr = R3 j02 (w) 1/R (8.119)
0 E−m
Putting these together

mR3 2
Z
1 1 1
q ∗ qdr = N 2 [2R3 j02 (w) (E − 1/R) + j (w) ] (8.120)
E−m E 0 E−mR
Therefore

2E(E − 1/R) + m/R


N −2 = R3 j02 (w) (8.121)
E(E − m)
which is the normalization term.

————————————————–
Problem 8.6: Given the wave function below, what is the normalization
term for state with arbitrary j?
!
Nκ jκ (pr)
ψjj3 κ (~r) = q Yjjl 3
4π −i E−M σ · r̂
E+m jκ−1 (pr)~

————————————————–
Bag Models 295

8.4.1 Magnetic Moments


Z X †
~ = (1/2)(~r × ~jem )d~r = (1/2)
µ dr~r × [ qi (~r)αi Qi qi (~r)]
bag
 
0 ~σ
αi =
~σ 0
Z R r
N2 X 2 (E − m)(E + m)
µ
~= Qi r dri j0 (wr/R)j1 (wr/R)
2 0 E2

×[(~r × ~σ )~σ · r̂ − ~σ · r̂(~r × ~σ )]


The square bracket as before, i.e., r and ~σ term separated is
Z R
2 2 w
µ= N ΣQi σi r3 j0 (wr/R)j1 (wr/R)( )dr
3 0 2E

(E−m)(E+m) w
where E = RE .
From an earlier step,
R w
R4
Z Z
3
r j0 (wr/R)j1 (wr/R)dr = 4 w3 j0 (w)j1 (w)dr
0 w 0

R4 w 3
= [ (3j12 (w) − j0 (w)j2 (w))] (8.122)
w4 4
now,
w
w3 2
Z
w3 j0 (w)j1 (w)dw = [3j1 (w) − j0 (w)j2 (w)] (8.123)
0 4

w3 sin w cos w 2 sin w 3 sin w sin w cos w


= [3( 2 − ) − ( 3
− − 3 2 )] (8.124)
4 w w w w w w
q
Using the boundary condition as earlier cos w = sinww − E−m
E+m sin w, and
p
with w = R (E − m)(E + m), we obtain:

w
w3
Z
1
w3 j0 (w)j1 (w)dw = j0 (w) [4E + 2m − 3/R] (8.125)
0 4 E−m
(Note that for m → 0 it reduces to the proper form obtained earlier)
Z
1
w3 j0 (w)j1 (w)dw = sin2 w(4w − 3) (8.126)
4
hence
296 Group Theory in Particle, Nuclear, and Hadron Physics

2 2X R4 w 3 1 w
µ= N Qi σi 4 j0 (w) (4E + 2m − 3/R) (8.127)
3 w 4 E−m RE
Substituting for the N 2 term , we obtain for k = −1,
R 4ER + 2mR − 3
µ=( ) eq (8.128)
6 2(ER)2 − 2ER + mR
By taking the values ER = 2.61, mR = 1.01, µ = 0.1675(eq R).
So,
eq e
(2mp R)
µp = 0.1675 (8.129)
e 2mp
e
P
where 2m p
is nuclear magneton, for the proton eq = e and R is in fermis.
Hence, µp = 1.59R.
So a similar problem as for the massless case presents itself here too.

8.4.2 Axial Vector Coupling Constant


We have seen that
gA 5 5
= < σz >= (1 − 2 < lz >) (8.130)
gv 3 3
And here we can write for the massive quarks in a bag,

R
E−m 2
Z
2 E+m
< σz >= N (( )j0 (wr/R) − 1/3( )Jk (wr/R))r2 dr (8.131)
0 E E

Then,

E(E − m) R3 j02 (w)


2 3
(2E(E − 1/R) + m/R)R j0 (w) 2(E − m)E

((E + m)(2E − 1/R) − 1/3(E − m)(2E − 3/R))

gA 2/3ER(ER + 2mR) − mR
= 5/3 (8.132)
gv 2ER(ER − 1) + mR
For m → 0 this reduces to
w
gA = 5/3( 3(w−1) ) as before.
gA
Take ER = 2.61 , mR = 1.01 to get gv = 1.25. Compare this to 1.01 of
the m → 0 limit.
Bag Models 297

8.4.3 Spin Structure of the Nucleon


We now derive the spin of the nucleon, as accomplished above for massless
quarks, but now with massive quarks. We require

u↑ =< p↑ |Pu (3)P ↑ (3)|p↑ > (8.133)


We need to calculate
Z R E+m 2 1 E−m 2
↑ 2 2
< ψ1/2,1/2 (3)|P (3)|ψ1/2,1/2 (3) >= N ( j0 (wr/R) + j1 (wr/R))r dr
0 E 3 E

E(E − m) E + m R3 2 2E − 1/R
= ( j (w)
R3 j02 (wr/R)(2E(E − 1/R) + m/R) E 2 0 E−m

1 E − m R3 2 2E − 3/R
+ j0 (w) )
3 E 2 E−m
E(4E/3 − 1/R)2mE/3
=
2E(E − 1/R) + m/R
and

R
2E−m 2
Z
↓ 2
< ψ1/2,1/2 (3)|P (3)|ψ1/2,1/2 (3) >= N r j1 (wr/r)dr
0 3 E

(E − m)(2E/3 − 1/R)
=
2E(E − 1/R) + m/R
So

1 (E − m)(2E/3 − 1/R)
u↑ =
3 2E(E − 1/R) + m/R

5 E(4E/3 − 1/R) + 2mE/3 2(E(11E − 9/R) + m(4E + 3/2R))


+ =
3 2E(E − 1/R) + m/R 9(2E(E − 1/R) + m/R)

u↑ = 2[ER(11ER − 9) + mR(4ER + 3/2)]/F (8.134)


where F = 9[2ER(ER − 1) + mR]
2mE
u↓ = E(4E/3 − 1/R) + + 5(E − m)(2E/3 − 1/R)
3

= 2[ER(7ER − 9) + mR(−4ER + 15/2)]/F


Next
2 (E − m)(2E/3 − 1/R)
d↑ =
3 2E(E − 1/R) + m/R
298 Group Theory in Particle, Nuclear, and Hadron Physics
1 E(4E/3 − 1/R) + 2mE/3
+ = [ER(8ER − 9) + 2mR(3 − ER)]/F
3 2E(E − 1/R) + m/R
and find then

d↓ = [ER(10ER − 9) + mR(2ER + 3)]/F

————————————————–
Problem 8.7: Check that u↑ + u↓ = 2 and d↑ + d↓ = 1.
————————————————–

Thus

5 (2/3)ER(ER + 2mR) − mR
g1p = ( ) (8.135)
18 2ER(ER − 1) + mR

g1n = 0 (8.136)
gA
Using gvfrom Equation 8.132, we see that these do satisfy the Bjorken
sum rule [66],
Z 1
gA
(g1p (x) − g1n (x))dx = /6 (8.137)
0 gv
In the non-relativistic limit mR → 0 the standard quark model results are
obtained. For the standard case m → 0 , we obtain from above,
11w0 − 9 ↓ 7w0 − 9
u↑ = , u =
9w0 − 9 9w0 − 9
8w 0 − 9 10w0−9
d↑ = , d↓ = (8.138)
18w0 − 18 18w0 − 18
where w0 = 2.04. This result is the same as what we obtained above for the
massless quarks directly in the MIT bag model.
Sehgal [69] defines the spin of the proton as the sum of total spin Sz and the
total angular momentum Lz of its partons. We take the complete integrated
function of these as < Sz > and < Lz >. In fact we used this expression in
the calculation of massless quarks in the MIT bag model for the quantity ggAv ,
the axial vector coupling constant. So as we found earlier,
9 p
Sz = (g + g1n )
5 1
1
− Sz
Lz = (8.139)
2
This tells how much spin resides in the quarks and how much in its orbital
angular momentum. In the MIT bag model its total integrated value, which
Bag Models 299
TABLE 8.2: Experiment [E142] compared to predictions of MIT bag
model (for massless quarks w0 = 2.04 and for m 6= 0, ER=2.61, and
mR=1.01); the mR → ∞ case corresponds to non-relativistic quark model;
best configuration mixed cases (1) and (2); and the deformed nucleon model
case
u↑ u↓ d↑ d↓ gA g1p g1n Sz
Expt - - - - 1.26 0.129 -0.031 0.176
(m = 0) 1.436 0.564 0.391 0.609 1.09 0.182 0 0.327
(m 6= 0) 1.499 0.501 0.375 0.625 1.247 0.208 0 0.374
(mR → ∞) 5/3 1/3 1/3 2/3 5/3 5/18 0 0.5
mixed (1) 1.66 0.34 0.35 0.65 1.59 0.271 +0.006 0.499
mixed (2) 1.65 0.35 0.34 0.66 1.614 0.273 +0.004 0.499
Def. nucl. 1.464 0.536 0.333 0.666 1.26 0.187 -0.0225 0.297

was presumed in the quark model to be the integration of the current quarks,
leads to corresponding values of the constituent quarks.
We display these results in Table 8.2 [case 1 - [70] , case 2 - [71]. For the
sake of comparison, we also show the results for the deformed nucleon case
and the best configuration-mixed quark model wave functions for the same
quantities.
In the non-relativistic quark model Sz = 0.5, this means that all the spin
in the proton/neutron sits in the quarks intrinsic spins only. The same is found
to be true for both the configuration mixed cases. This is at variance with the
experimental values and hence these models fail in this regard. In the MIT
bag model with m = 0, 65% of proton spin comes from the spin of the quarks
and the rest comes from the orbital angular momentum between the quarks.
The corresponding nucleon for m 6= 0 case is 74% and 26%, respectively. For
the deformed nucleon case 59% of the spin of proton sets in the intrinsic spin
of the quarks and 41% in the orbital angular momentum degrees of freedom
of the quarks. Thus the resolution of the EMC effected of the “spin crisis”
in these pictures: in the MIT bag model and the deformed nucleon picture
comes entirely in terms of the orbital angular momentum degrees of freedom.
No strange quarks contribute here, as there are no strange quarks present in
proton and neutron.
So does it means that the MIT bag model is a good model to describe the
spin structures of the nucleons? But note that it may not be too bad for g1p ,
but it fails to describe the negative and non-zero value of g1n . In the MIT bag
model g1n = 0 and thus it is failing in the producing this quantity. This shows
that partial and piecemeal successes of a particular model should be taken
with a pinch of salt.
The configuration mixed quark models also fail to reproduces g1n . From
the Table 8.2, we see that these are positive and quite small in magnitude
which does not match the negative −0.003 value. Hence these models fail in
this regard too.
300 Group Theory in Particle, Nuclear, and Hadron Physics

Only the deformed nucleon model is found to give consistent and good
description of all the quantities here.
Note that the MIT bag model and the configuration mixed quark models
gave pretty good descriptions of non-polarized physical quantities. And this
was the primary reason of their appeal to the physicists. However these models
find their Waterloo in terms of the polarized quantities as discussed here.
Surprisingly, the deformed nucleon picture continues to be successful even in
describing the polarized observables, as shown here.
Note that as discussed in detail in the next chapter in studying the re-
lationship between the constituent quarks and the current quarks, we show
that
Z x
u↑ = u↑ (x)dx (8.140)
0
Though the gluons have not been included explicitly in the MIT bag model,
the integration over x on the complete 0 to 1 range ensures that u↑ is some
genuine quasi-quark value where implicitly all the sea of quark-antiquark and
gluons contribute to the current or valence quarks.

8.5 Scalar and Vector Confining Potentials


Whatever potential that we take in confining relativistic quarks,these must
obey the symmetry imposed by the Lorentz transformation. This is a very
general restriction and here we demand (in generalizing the static spherical
bag and the MIT bag model cases) that these central potentials be of scalar
and vector kinds. We take the vector component of the confining potential as
the zeroth-component of the four vector. We thus bracket the scalar potential
U (r) with the mass m, and the zeroth component of the vector potential V0 (x)
goes with the energy E in the Dirac equation.
So let the quarks be confined in a central potential U (r) so that they are
represented by single potential wave function satisfying the Dirac equation.

[γ0 E + i~r · ~δ − (m + U (r))]ψ(~r) = 0 (8.141)


where U (r) is given in a general form,

U (r) = 1/2(1 + aγ0 )M (r) (8.142)


The parameter a can in principle be allowed to take any value [72]. Of
special interest are the cases:
(a) a = 1; which leads to the exact SU (2) symmetry and hence to spin-
orbit doublet degeneracy. It was studied by Bell and Ruegg in the context of
the Melsh transformation [73].
Bag Models 301

(b) a = 0; which is the scalar potential case of the MIT bag model (already
studied here) and the Dubna bag model.
Using [74]:

∞ r>R
M (r) = (8.143)
0 r<R
We take M(r) in a general form as

M (r) = cn rn (n = 0, 1, 2, 3, )˙ (8.144)
As we saw earlier,
 
N ig−1 (r)/r
ψ(~r) = √ χ1/2 (8.145)
4π i~σ · r̂f−1 (r)
where
Z 1
−2 2 2
N = (g−1 (r) + f−1 (r))r2 dr (8.146)
0

Note that the detailed form of the upper g−1 (r) and the lower j−1 (r)
components of the quark spinor will depend on the potentials chosen. For
example, as we saw, these are related to the Spherical Bessel function j0 (r)
and j1 (r) in the MIT bag model. In most cases it is not possible to obtain
them analytically. And this is also true of the general potential above.
We saw earlier that

R
RR 2
5N 2 4/3 r j−1 (r)dr
Z
gA
= (r2 g−1
2
(r) − 1/3r2 j−1
2
(r))dr = 5/3[1 − R R 0 ]
gv 3 0 r 2 (g 2 + j 2 )dr
0 −1 −1
(8.147)
As a check in the MIT bag model case we found,

R R
R3 sin2 w(2w − 3)
Z Z
r2 j−1 (r)dr = r2 j1 (wr/R)dr = (8.148)
0 0 2 w3
and
R
w3
Z
r2 (g−1
2 2
(r) + j−1 (r))dr = (8.149)
0 2R3 (w − 1) sin2 w
And so
gA 2w − 3 w
= 5/3[1 − ] = 5/3 (8.150)
gv 3(w − 1) 3(w − 1)
which is good.
Define
302 Group Theory in Particle, Nuclear, and Hadron Physics

RR
0
r2 f−1
2
(r)dr
X = RR (8.151)
2 (r) + f 2 (r))dr
r2 (g−1
0 −1

Thus gA = 5/3[1 − 4/3X].


We shall express all relevant quantities in terms of X here as we have to
calculate them numerically for these potentials:

Z R Z R
↑ 2 1
< ψ1/2,1/2 (3)|P (3)|ψ1/2,1/2 (3) >= N ( r2 g−1
2
dr + r2 j−1
2
dr)
0 3 0

= 1 − 2/3X
This is also equal to < ψ1/2,1/2 (3)|P ↓ (3)|ψ1/2,1/2 (3) >. Also,

< ψ1/2,1/2 (3)|P ↓ (3)|ψ1/2,1/2 (3 >=< ψ1/2,−1/2 (3)|P ↑ (3)|ψ1/2,−1/2 (3) >= 2/3X

Finally

u↑ =< p↑ |Pu (3)P ↑ (3)|p↑ >= 5/3 − 8/9X (8.152)


and ultimately we obtain

u↓ = 1/3 + 8/9X
d↑ = 1/3 + 2/9X
d↓ = 2/3 − 2/9X (8.153)
Thus

g1p = 5(1 − 4/3X)/18


g1n = 0 (8.154)
From the Bjorken sum rule,
Z 1
gA
(g1p (x) − g1n (x))dx = ( )/6 (8.155)
0 gv
Find g1p = ( ggAv )/6
This result is independent of any details of the above potential and is a
generic result for any scalar-vector confining potential. So fixing X to get
gA p
gv = 1.26 (which is the experimental value), it gives g1 = 0.21 in these
models. This is much larger than the EMC result of 0.14, and also larger than
the Eliss-Jaffe sum rule result of 0.187 (to be derived in the next chapter).
So it is quite clear that in all these models (including the MIT bag model),
it is not possible to simultaneously fit both gA and g1p . Success in terms of
fitting any one of these guarantees failure in terms of the other one for all
Bag Models 303

these models. And also all of these models yield g1n = 0. This is a gross failure
with respect to the experimental value of -0031 [57].
The values of u↑ , u↓ , d↑ and d↓ , g1p and gA for various potentials given
above could be evaluated numerically by calculating X. These are listed in
[72]. These scalar-vector potentials, which work so well for unpolarized phys-
ical observables, fail (as does the MIT bag model also), in the case of the
polarized physical observables. Thus, polarized physical observables are more
discriminating and demanding and therefore should be the final arbiter in
deciding as to which models should be discarded and which ones be retained.
In all these different ways of modelling QCD to understand confined
quarks, two basic aspects come into play. The first one is as to how well the
symmetry structure of QCD gets included in a particular model. The second
one pertains to the phenomenological manner in which the non-perturbative
aspect of confinement is incorporated. In the above models, the difference in
the non-perturbative aspects arise from the difference in the calculated value
of ‘X’. As the exact value of it is has no consequence for the conclusions ar-
rived herein, the non-perturbative aspects of QCD in these models are washed
away. Thus one may conclude that actually these models are not taking proper
account of the symmetry structure of QCD. This is a general and basic con-
clusion we can draw here. On the same grounds, therefore, the success of the
deformed nucleon model demonstrates that it seems to incorporate both the
symmetry structure of QCD and its non-perturbative aspects in a rather more
faithful manner.

8.6 Solutions of Problems

Solution 8.1:
 
~ = ~r × p~, [H, L]
~ = [~ ~ 0 ~σ
[xi , pj ] = iδij , L α · p~ + βm, L], α
~ = ,
  ~σ 0
1 0
β= .
0 −1
As β is diagonal, it commutes with L. ~

[αx px + αy py + αz pz , Lz ], ( with Lz = xpy − ypx )

= αx [px , xpy − ypx ] + αy [py , xpy − ypx ] + αz [pz , xpy − ypx ] = αx [px , x]py −
αy [py , y]px = −iαx py + iαy px = −i(~ α × p~)z
Therefor [~ ~
α · p~, L] = −i~
α × p~  
So L is not conserved. Take Σ ~ = ~σ 0 .
0 ~σ
304 Group Theory in Particle, Nuclear, and Hadron Physics
~ = [~
[H, Σ] ~
α · p~, Σ]
z-component [~ α · p~, Σz ]
     
0 ~σ · p~ σz 0 σ 0 0 ~σ · p~
− z
~σ · p~ 0 0 σz 0 σz ~σ · p~ 0

   
0 ~σ · p~σz 0 σz ~σ · p~
= −
~σ · p~σz 0 σz ~σ · p~ 0

 
0 ~σ · p~σz − σz ~σ · p~
=
~σ · p~σz − σz ~σ · p~ 0

~σ · p~σz = σx px σz + σy pz σz + σz pz σz = −iσy px + iσx py + pz

σz ~σ · p~ = σz σx px + σz σy py + σz σz pz = iσy px − iσz py + pz

   
0 2i(σx py − σy px ) 0 (σ × p)z
[H, Σz ] = = 2i
2i(σx py − σy px ) 0 (σ × p)z 0

~ = 2i~
[H, Σ] α × p~
Therefore, J~ = ~l + (1/2)Σ
~ is conserved. [H, J]
~ = 0.
~
The eigenvalue of (1/2)Σ are ±(1/2).
————————————————–
Solution 8.2:
   
0 1
g−1 (~
r )Y0 (r̂)
0  
P ↑ ψj=1/2,j3 =1/2 = 


1 
q
if−1 (~r)(− 13 Y10 (r̂)

)
0
 
0
P ↓ ψj=1/2,j3 =1/2 = 
 
0 
q
if−1 (~r) 23 Y11 (r̂)
1
 
0
P ↑ ψj=1/2,j3 =−1/2 = 
 
1 
q
if−1 (~r)(− 23 Y1−1 (r̂) )
0
   
0 0
 −1 g (~
r )Y0 (r̂)
↓ 1 
P ψj=1/2,j3 =−1/2 =  
0 
q
if−1 (~r) 13 Y10 (r̂)

1
————————————————–
Bag Models 305

Solution 8.3:
2
This relation holds because (~σ · r̂) = 1. So the eigenvalues of ~σ · r̂ must
be +1 or −1. Let us choose r̂ parallel to x3 axis and θ = 0. In the spherical
harmonics one can confirm that the eigenvalue is −1 and independent of j and
l.
————————————————–
Solution 8.4:

ηµ T µν = (i/2)(ψ̄ηµ ψ µ ∂ ν ψ − ∂ ν ψ̄ηµ γ µ ψ)θV − ηµ g µν L


The first bracket term,

(1/2)(iψ̄γ µ ηµ ∂ ν ψ − ∂ ν ψ̄iγ µ ηµ ψ)θV


= (1/2)(−ψ̄∂ ν ψ − (∂ ν ψ̄)ψ)θV = −(1/2)∂ ν (ψ̄ψ)θV = −Bη ν θV
(using Bη ν = (1/2)∂ ν (ψ̄ψ) on the surface).
and

−η µ g µν L = −η ν L = −η ν [[(i/2)ψ̄γ µ ∂ µ ψ − ∂ µ ψ̄γ µ ψ − B]θν − (1/2)ψ̄ψδs ]

= −(i/2)(ψ̄γ µ ∂ µ ψ − δ µ ψ̄γ µ ψ)η ν θV + η ν BθV + (1/2)η ν ψ̄ψδs


The first bracket term and the last term are zero on the bag surface. And
so ηµ T µν = 0. Thus no energy escapes from the boundary of the MIT bag
surface.
————————————————–
Solution 8.5:
Now,

        
0 1 0 0 ~σ 0 1 1 0 ~σ 0 ~σ 0
γ ~γ γ5 = = =
0 −1 −~σ 0 1 0 0 −1 0 −~σ 0 ~σ

In bag
Z
d~rq † (~r)γ 0~γ γ5 q(~r)

N2
Z   
~σ 0 j0
= d~r(j0 , −i~σ · r̂j1 )
4π bag 0 ~σ i~σ · r̂j1
N2
Z Z
= drr2 dr̂(j02~σ + ~σ · r̂~σ~σ · r̂j12 )
4π bag
now
1
~σ · ~r~σ~σ · ~r = [~σ · ~r~σ~σ~r]
r2
306 Group Theory in Particle, Nuclear, and Hadron Physics

~σ · ~r~σ~σ · ~r = (xσx + yσy + zσz )(iσx + jσy + kσz )(xσx + yσy + zσz )

= [i(x−iyσz +izσy )+j(ixσz +y −izσx )+k(−ixσy +iyσz )]×(xσx +yσy +zσz )

Take i part

(x − iyσz + izσy )(xσx + yσy + zσz ) = i[2x(xσx + yσy + zσz ) − (x2 + y 2 + z 2 )σx ]

~ − r 2 σx ]
= i[2x(~r · σ)
Putting in all, we get
1 1
2
[~σ · ~r~σ~σ · ~r] = 2 [2~r(~r · ~σ ) − r2~σ ]
r r
2
R
~r~r here gives 1/3r , the scalar part. In the d~r integration only this will
survive. Thus it separates out the r~σ term.
→ (1/r2 )[(2/3)r2~σ − ~r · ~σ ] = −(1/3) r12 (r2~σ ) = −(1/3)~σ
Z Z

d~r[~σ · ~r~σ~σ · ~r] = −(1/3)~σ d~r = − ~σ
3
Putting it above

Z Z R 3
~ r)|n >= N 2
X ~τ
dr < p|A(~ drr2 (j02 (wr/R)−(1/3)j12 (wr/R)) < p| ~σ |n >
0 i=1
2

next Z R
2
N drr2 (j0 − (1/3)j1 )
0
Z R Z R
2
=N ( drr2 j02 (wr/R) + r2 j12 (wr/R) 2
− N (4/3) drr2 j12 (wr/R))
0 0
Z R
2
=1−N drr2 j12 (wr/R))
0

R
R3 2 sin2 w
Z
sin w
drr2 j12 (wr/R) = 2
(− +1+ cos w)
0 2w w2 w
using j0 (w) = j1 (w)
R
R3 sin2 w
Z
drr2 j12 (wr/R) = (2 − 3/w)
0 2 w2
Bag Models 307

so Z R
N2 drr2 (j0 − (1/3)j1 ) = 1 − (1/3)(w/(w − 1))
0
Note that,
X ~τ
< p| ~σ |n >= 5/3
i
2
This is the axial vector coupling constant of the quark model. And hence,
Z
~ r)|n >= (1 − (1/3)(w/(w − 1))(5/3) = 1.08
d3 r < p|A(~

This works well with respect to the experimental value. Note that the
relativistic correction of 0.65 to the non-relativistic model of 5/3 takes place
at the right direction as having the value of ggAv .
————————————————–
Solution 8.6:

R
N2 E−m 2 2
Z Z
∗ l∗
ψjj ψ
3 κ jj3 κ
d~r = κ [jκ2 (pr) + ( ) jκ−1 (pr)]r2 dr(Yjj l
Yjj )
4π 0 E+m 3 3

Now,

l ∗
l 1
Yjj Yjj = [(j + j3 ) + (j − j3 )] = 1
3 3
2j
j = l + 1/2 , κ = −(j + 1/2) = −l − 1
jκ = j−l−1 = (−1)l+1 jl+1−1 = −(−1)l jl
jκ2 = jl2
jκ−1 = j−(l+2) = (−1)l+2 jl+2−1 = (−1)l jl+1
2 2
jκ−1 = jl+1

R R
E−m
Z Z Z

ψjj ψ
3 κ jj3 κ
d~r = Nκ2 [ jκ2 (pr)r2 dr + 2
jk+1 (pr)r2 dr]
0 E+m 0

Take pr = wr/R,
3 Z w
R3 w 2
Z Z
∗ 2 R 2 2
ψjj3 k ψjj3 k dr = Nk [ j (w)w dw + 3 j (w)w2 dw]
w 0 l w 0 l+1

Nκ2 R3 E − m X3 2
[(1/2)w3 (j22 (w)−jl−1 (w)jl+1 (w))+ (j (w)−jl (w)jl+2 (w))]
w 3 E + m 2 l+1
Nk2 R3 4l + 4 2m
= [E(2jl2 (w) + 2jl+1
2
(w) − jl (w)jl+1 (w)) + jl (w)jl+1 (w)]
2(E + m) w w
308 Group Theory in Particle, Nuclear, and Hadron Physics

The boundary condition is


s s s
E−m Nk2 R3 E−m E − m 4l + 4 2m E−m
jl (w) = jl+1 (w) = [E(2 + 2 − )+ ]
E+m 2(E + m) E+m E+m w w E+m

p
Putting X = (E − m)(E + m)R,

Nκ2 R3 2E 2 R + m − 2E(l + 1)
Z

ψjj ψ
3 κ jj3 κ
d~r = 2jl2 (w)[ ]
2(E + m) (E − m)R
2E(ER − −(l + 1)) + m
Nκ−2 = R2 jl (w)[
]
(E + m)(E − m)
————————————————–
Solution 8.7:

2
u = u↑ + u↓ = (E(11E − 9/R)
9(2E(E − 1/R) + m/R)
+m(4E + 3/2R) + E(7E − 9/R) + m(−4E + 15/2R)) = 2
E(8E − 9/R + 10E − 9/R) + m(−2E + 6/R + 2E + 3/R)
d = d↑ + d↓ = =1
9(2E(E − 1/R) + m/R)
————————————————–
Chapter 9
Hadron Physics

9.1 Harmonic Oscillator Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309


9.2 Configuration Mixing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
9.3 Deformed Nucleon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323
9.4 Weak Charges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 325
9.5 Magnetic Moments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333
9.6 Spin of Nucleon in the Quark Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 339
9.6.1 Spin of a Deformed Nucleon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 340
9.6.2 Spin of Nucleon with Configuration Mixed Wave
Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348
9.7 Colour Confinement in QCD and Deformed Baryons . . . . . . . . . . . 356
9.8 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 359

In Chapter 7 we had utilized the wave function of the group SU (6)SF ⊃


SU (3)F ⊗ SU (2)S to perform calculations in the non-relativistic quark model.
The same wave function was also used in the bag model in Chapter 8.
Here we shall expand the above quark model group structure to allow for
orbital excitations between quarks. This larger group is SU (6)SF ⊗ SO(3)l ⊃
SU (3)F ⊗ SU (2)S ⊗ SO(3)l where SO(3)l is the 3-dimensional rotation group.
This is achieved through the harmonic oscillator model.

9.1 Harmonic Oscillator Model


The classical harmonic oscillator
p potential in one dimension is V (x) =
1
2 Kx
2
= 12 mω 2 x2 (where ω = k/m). Classically the total energy is continu-
ous and hence any real number value for ω is possible. Quantum mechanically
however the energy is quantized as En = ~ω (n=0,1,2...).
The deep-rooted significance of the oscillator model in physics is due to
the fact that within this model a transition can be performed from a contin-
uous basis to a discrete basis and vice-versa. Hence the harmonic oscillator is
the essential bridge which connects the continuous (classical) to the discrete
(quantum). Thus it is the harmonic oscillator which allows the classical (con-

309
310 Group Theory in Particle, Nuclear, and Hadron Physics

tinuous) and the quantum (discrete) realities to coexist peacefully, thereby


providing an underlying Manichaean duality to the same physical reality.
In the 3-dimensional Cartesian coordinate system we write the harmonic
oscillator potential as
1
U (~r) = mω 2~r2
2
1
mω 2 x2 + y 2 + z 2

= (9.1)
2
Thus there are three independent harmonic oscillators and one obtains the
energy eigenvalue as [44]

 
3
Enx ,ny ,nz = hω nx + ny + nz + (nx , ny , nz = 0, 1, 2.....) (9.2)
2
7
where N = nx + ny + nz . For N = 2, E = 2 hω. This corresponds to six
degenerate states given as (nx , ny , nz ):

(2, 0, 0) , (0, 2, 0) , (0, 0, 2) , (1, 1, 0) , (1, 0, 1) , (0, 1, 1) (9.3)


This is puzzling as in the SO(3)l model the l = 2 states are only fivefold
degenerate (ml = −2, −1, 0, 1, 2). How do we understand this extra degeneracy
term here?
A transformation to the polar coordinate system here assists in further
comprehension. In that system we write the harmonic oscillator Hamiltonian
as:

p~2 1
H= + mω 2~r2 (9.4)
2m 2
 
The Schroedinger equation is p~ = −i~∇ ~ ,
 
1 h ~2 mω 2
Hψ = ~ω ∇ + ~r ψ = Eψ (9.5)
2 mω ~
q q
Replacing ~r0 = ~r/ mω~ ~0 =∇
,∇ ~ ~
mω yields,
r
1 n ~0 2 ~0 2 o 0 h
H = ~ω ∇ + r r = r/b, b = (9.6)
2 mω
Here b plays the role of the characteristic length of the oscillator or the
scale of the harmonic oscillator potential.

mω 1
= 2 = α2
ν= (9.7)
h b
Thus Schroedinger equation is separable into two equations in terms of the
Hadron Physics 311

radial and the angular coordinates:

ψnlm = Rn,l Yl,ml (θ, φ) (9.8)


l is the angular momentum quantum number, ml is its z-projection and n is
the number of radial nodes between r = 0 and r = ∞. Then we obtain,
 
3
En = N + ~ω (9.9)
2
and
 
3
Enl = 2n + l + ~ω (9.10)
2
Note N = 2n + l = nx + nx + ny . The above is the convention we uti-
lize here. There is another convention which is often used. Therein En0 l =
2n0 + l − 12 hω = N 0 + 23 hω, N 0 = 0, 1, 2 · · · . Then N 0 = 2 (n0 − 1) + l
 

whence (n0 − 1) represents the number of nodes.


One obtains the wave function as [44],
l+ 12 1 2 2
ψN lm = N (αr)l Ln (α2 r2 )e− 2 α r
Ylm (θ, φ) (9.11)
l+ 12
where Ln (α2 r2 ) is Laguerre polynomial [26],
n 
n + l + 12 (−t)ν

l+ 21
X
Ln (t) = (9.12)
n−ν ν!
ν=0

The normalization factor is


2
N = √ 2α3 n!
(9.13)
π(n + l + 12 )(n + l − 12 ) · · · 32 . 12
A few lower states ψN l are given below

4α3 1 1 2 2
ψ0s = ( √ ) 2 e− 2 α r Y00 (Ω)
π
r
2 4α3 1 1 2 2
ψ0p = ( √ ) 2 αre− 2 α r Y1m (Ω)
3 π
r
4 4α3 1 2 2 − 1 α2 r2 m
ψ0d = ( √ )2 α r e 2 Y2 (Ω)
15 π
r
2 4α3 1 3
 
1 2 2
ψ1s = ( √ )2 − α2 r2 e− 2 α r Y00 (Ω) (9.14)
3 π 2
Note that the radial part of these functions for small values of r behaves as
312 Group Theory in Particle, Nuclear, and Hadron Physics

_3_
A

2P 3

FIGURE 9.1: Jacobi coordinates for three-dimensional space

rl , i.e., it vanishes at the origin except for the s-wave (l = 0) case. Thus with
respect to its behaviour at r = 0, the s-state is completely distinct from the
other states of the harmonic oscillator potential. How this effect may manifest
itself physically shall be discussed in the last section of this chapter.

————————————————–
Problem 9.1: Check the normalization of the above state ψ0d .
————————————————–

Let us now imagine three confined quarks of mass m each, interacting via
a harmonic oscillator potential and located at r~j (j=1,2,3) as
3
X p~i 2 1 X ~2
H= + K rij (9.15)
i=1
2m 2 i<j

where r~ij = r~i − r~j

The definition of the Jacobi coordinates is provided in Figure 9.1.

1 1 ~ = 1 (r~1 + r~2 + r~3 ) (9.16)


~ = √ (r~1 − r~2 ), ~λ = √ (r~1 + r~2 − 2r~3 ), RCM
ρ
2 6 3

where
Hadron Physics 313

d~
ρ d~λ ~
dRCM
p~ρ = m
, p~λ = m , and pCM
~ =M (9.17)
dt dt dt
where M = 3m. One finds,

~ 2
pCM
H= + H0 (9.18)
2M
where

p~ρ 2 p~λ 2 3  2 ~ 2
H0 = + + K ρ
~ +λ (9.19)
2m 2m 2
Note we define,
r
3k 1 √ 1
ω= , α = (3km) 4 = mω, ν = mω = 2 = α2 (9.20)
m b

————————————————–
Problem 9.2: Show that, given the Hamiltonian Equations 9.15 and 9.18
(15) hold true.
————————————————–

The 3-dimensional harmonic oscillator states correspond to uncoupled


ρ and λ equations. The total angular momentum is L ~ =L ~ρ + L ~ λ for bound
states. The total space wave is the product of the space wave function of λ
and ρ oscillators. Hence parity of the bound states is P = (−1)lρ +lλ . The
total angular momentum of the bound state is J~ = L
~ +S ~ where S ~ is the total
quark spin. The bound states can be labelled by the eigenvalue lρ and l~λ and
~
by the number of radial excitations nρ and nλ (nρ,λ = 0, 1, 2 · · · ) of the two
oscillators. The principal quantum number is

N = 2nρ + 2nλ + lρ + lλ (N = 0, 1, 2...) (9.21)


p
The total bound states mass is M = M0 + ωN ; ω = 3K/ω with M =
3mq + 3ω + V0 (note mq = m and V0 is a constant). So far all the states
belonging to the same principal quantum numbers N are degenerate.
Now we look at the wave function of this harmonic oscillator. For example
for the ρ-oscillator,
ml ρ
ψnρ lρ (ρ) = Rnρ lρ Ylρ (ρ) (9.22)
For the total wave function we use the notation,
N
ψlmΣ {Σ = S, A, ρ, λ} (9.23)
These are listed in Table 9.1. Hence Σ defines the overall symmetry of the
orbital wave function of SO(3)l . Thus the ground state wave function being
314 Group Theory in Particle, Nuclear, and Hadron Physics
TABLE 9.1: Baryon space wave functions in the Harmonic Oscillator
Model – these are used in Table 9.2.
α2 = mω, N = 2nλ + 2nρ + lλ + lρ , n = nλ + nρ , l = lλ + lρ , ~l = l~1 + l~2 .
α2
(ρ2 +λ2 )
X = e− 2

N
N n l ψlmΣ (Σ = S, A, ρ, λ) lP SU (6)F S
3
0 0 0 0
ψ00S : α3 X 0+ 56
q 4π 2
1 0 1 1
ψ1mλ : 83 απ λYm1 (Ωλ )X 1− 70
q 4
1 0 1 ψ1mρ1
: 83 απ ρYm1 (Ωρ )X 1− 70
q 3
2 1 0 2
ψ00S : 13 α3 [3 − α2 (ρ2 + λ2 )]X 0+ 56
q π2
2 8 α5 2 2 2 2
2 0 2 ψ2mS : 15 π q Ym (Ωρ ) + λ Ym (Ωλ )]X
[ρ 2+ 56
5
2 1 0 ψ00λ2
: 13 α3 (ρ2 − λ2 )X 0+ 70
5
π 2
2
: − 38 α1 ρλ µµ0 Cµµ 110 1 1 +
P
2 0 2 ψ00ρ 0 0 Yµ (Ωρ )Yµ0 (Ωλ )X 0 70
π
q 2
2 8 α5 2 2 2 2
2 0 2 ψ2mλ : 15 π [ρ Ym (Ωρ ) − λ Ym (Ωλ )X 2+ 70
5
2
: 83 α1 ρλ µµ0 Cµµ 112 1 1 +
P
2 0 2 ψ2mρ 0 m Yµ (Ωρ )Yµ0 (Ωλ )X 2 70
π 25
2
: 38 α1 ρλ µµ0 Cµµ 111 1 1 +
P
2 0 2 ψ1mA 0 m Yµ (Ωρ )Yµ0 (Ωλ )X 1 20
π2

N = 0, L = 0+ should be given as

0 0 0
ψ00S ∼ ψ00 (ρ)ψ00 (λ)

which properly normalized is (see Table 9.1)

α3 −α2 (ρ2 +λ2 )/2


0
e ψ00S = (9.24)
π 3/2
This is a totally symmetric wave function as

ρ2 + λ2 = 13 (r~12
2 + r~2 + r~2 )
13 23

Let us check the normalization of the above ground state wave function,

ψ ? ψd3 ρd3 λ = 1
RR
We know
R 2π Rπ
given φ=0
dφ θ=0
sinθdθ = 4π
 2 hR i
3 ∞ 2 2 R∞ 2
λ2
= α
π 32
(4π)2 0
ρ2 e−α ρ
dρ 0
λ2 e−α dλ

Using
Hadron Physics 315


1 Γ(n + 21 )
Z
2
e−px x2n dx = (9.25)
0 2 p(n+ 12 )
R∞ 3 √
2
λ2 1 Γ( 2 ) 1 π
Thus 0
λ2 e−α dλ = 2 (α2 ) 32 = 4 α3
 2 √ 2
α3 π
So π 3/2
(4π)2 ( 14 α3 ) =1

————————————————–
2
Problem 9.3: From Table 9.1 check the normalization of ψ2ms .
————————————————–
Unsolved Problem 9.1: From Table 9.1 check the normalization of the
2
state ψ00S .
————————————————–

In SU (6)SF groups we know that

6 ⊗ 6 ⊗ 6 = 56s ⊕ 70λ ⊕ 70ρ ⊕ 20A (9.26)


Where as per the group decomposition SU (3)F ⊗ SU (2)S these are,

56s = (10, 4) ⊕ (8, 2)

70 = (10, 2) ⊕ (8, 4) ⊕ (8, 2) ⊕ (1, 2)

20A = (8, 2) ⊕ (1, 4) (9.27)


The lowest (0~ω) state is unique and exhibits an (0s)3 configuration with
[3] symmetry and angular momentum l = 0. The next (1hω) has (0s)2 (0p)
with either [3] or [21] symmetry with l = 1 (in both the cases). Having excited
only one quark, one cannot build totally antisymmetric states. The states
with the [3] symmetry is the spurious one on account of the centre of mass
excitation and should be excluded.
For example, if we take ψ0 (r~1 , r~2 , r~3 ) as the (0s)3 state which is fully sym-
metric in the coordinates (r~1 , r~2 , r~3 ) with respect to the centre of mass for the
three quarks. Let us excite one quark to the p-state which can be described
as r~1 , ψ, r~2 ψ0 and r~3 ψ0 . Then the [21] mixed symmetric states are
1
ψρ = √ (r~1 − r~2 )ψ0 (9.28)
2
1
ψλ = √ (r~1 + r~2 − 2r~3 )ψ0 (9.29)
2
And [3] is
316 Group Theory in Particle, Nuclear, and Hadron Physics

1
ψS = √ (r~1 + r~2 + r~3 )ψ0 (9.30)
2
We can now remove the spurious symmetric states of centre mass motion
by taking the centre of mass coordinate to be the origin,

r~1 + r~2 + r~3 = 0 (9.31)


Otherwise we take appropriate states or combinations of states to have an
appropriate symmetry.
Thus the first excited states N = 1 and Lπ = 1− having mixed symmetry
are
ρ
ψn=1,l=1 = ψ01 (ρ)ψ00 (λ) (9.32)
λ
ψn=1,l=1 = ψ00 (ρ)ψ01 (λ) (9.33)
The properly normalized N = 1, L = 1 are given in Table 9.1.
Next the second excited state has the configuration of the 2~ω excitation:
2 2 2
(0s) (0p) , (0s) (1s) and (0s) (0d). All these symmetries [3], [21], and [111]
are possible for the first case. For the second and the third one only [3] and [21]
symmetry is possible while [111] is excluded. There are five spurious states here
– two coming from the 0~ω internal excitations and three from the 1~ω inter-
2 2
natal excitations. In the harmonic oscillator potential (0s) (0d) , (0s) (1s)
2 ~ coordinate in any
and (0s) (0p) are degenerate. The presence of the RCM
of these indicates spurious components. These are eliminated by appropriate
linear combinations. Thus the trick is to take such linear combinations which
drop out any RCM ~ dependence.
We provide one example here, that of the symmetric representation L =
0+ . Using the orbital states for each quark we determine,
√  
2 2 9 2 2 2 2
ψ (0s) (1s) ∼ − α (r1 + r2 + r3 ) ψ0
3 2
2 2 2
ψ (0s) (0p) ∼ α (r~1 · r~2 + r~1 · r~3 + r~2 · r~3 )ψ0 (9.34)
3
Finally we find,
√  
2 2 9 2 2 2 2
ψ (0s) (1s) ∼ − α (3RCM + ρ + λ ) ψ0
3 2
 
2 2 2 2 1 2 2
ψ (0s) (0p) ∼ α 3RCM − (ρ + λ ) ψ0 (9.35)
3 2

————————————————–
Unsolved Problem 9.2: Confirm the expression given in Equations 9.34
and 9.35.
Hadron Physics 317
TABLE 9.2: Symmetric SU (6)F S wave functions for N and
∆. Octet and decuptet only for spins, J = 21 and 32 respectively.
χS = χ( 32 , S3 ), χλ,ρ = χ( 21 , S3 )λ,ρ , φΣ(Σ=S,A,ρ,λ) , ψlmΣ
N
from
Table 9.1
N SU(6) lP SU (3)F octet SU (3)F decuptet
0 56 0+ √1 (χρ φρ + χλ φλ )ψ00S0 0
χS φS ψ00S
2
2 56 0+ √1 (χρ φρ + χλ φλ )ψ00S2 2
χS φS ψ00S
2
2 56 2+ - 2
φS [χS ψ2mS ]J= 3
2
2 70 0+ √1 (φλ [χρ ψ 2 − χλ ψ 2 ] -
2 00ρ 00λ
2 2
+φρ [χλ ψ00ρ + χρ ψ00λ ])
1
2 70 2+ √ (φρ [χS ψ
2
2
]
2mρ J= 1 φS √12 ([χρ ψ2mρ
2
]J= 3
2 2
2 2
+φλ [χS ψ2mλ ]J= 1 ) +[χλ ψ2mλ ]J= 3 )
2 2
2 20 1+ √1 (φλ [χρ ψ 2
2 1mA ]J= 12 -
2
−φρ [χλ ψ1mA ]J= 1 )
2

————————————————–

Note that these states have explicit RCM dependence. We eliminate RCM
by taking a suitable linear combination of the two as
r r
2 2 1 2
ψ(L = o+ ) = (0s) (1s) + (0s) (0p)
3 3
3
3 − α2 (ρ2 + λ2 ) ψ0

∼ (9.36)
2
which suitably normalized is listed in Table 9.1. All the other states of SO(3)l
up to N = 2 are listed in [58].
Now we have all the symmetric, mixed symmetric and antisymmetric states
of SO(3)l . For the groups SO(6)SF ⊗ SO(3)l all the fully symmetric states
of the three quark baryons have to be constructed. The antisymmetry of the
total system would arise from the antisymmetry in the orthogonal colour space
SU (3)C . These fully symmetric states of SU (6)SF ⊗ SO(3)l are listed here in
Table 9.2 for N and ∆ states.

9.2 Configuration Mixing


If the harmonic oscillator is an exact symmetry between quarks in hadrons,
then we expect that all the states in N = 0, N = 1, N = 2, etc. would be
degenerate in the N -shells with a separation of 1~ω excitation between them.
318 Group Theory in Particle, Nuclear, and Hadron Physics

But even in the ground state N = 0, N and ∆ are not degenerate in energy,
with ∆ being ∼ 300 MeV higher than the nucleon N . The same situation
holds for N = 1, N = 2, etc. higher excitation levels. So there have to be
some symmetry breaking terms present which, we hope, may be calculable
within the perturbative theory. It turns out that in the quark model this
simple expectation holds good allowing for a fair degree of success.
In nuclear physics we have several nucleons interacting with each other,
but due to specific nuclear properties we need not worry about multinucleon
interactions, whence a two-body interaction term is found to be successful
in describing the nucleus. In a baryon three-body force, entities like a three-
gluon vertex arise in QCD itself. These are definitely not negligible. But note
that in the non-relativistic quark model we may have to deal with putative
three-body forces between three constituent quarks, which have absorbed the
fundamental gluons and the sea of quark-antiquarks in it. The dominance of a
two-body N −N force is evident from phenomenological studies of the nucleus.
We presume that the same holds of hadrons in terms of the forces between
quarks as well.
In fact when we use a harmonic oscillator potential as above, it is indeed
an acknowledgement of the above assumption. How does it fit into a more
general interaction between quarks in hadrons? So take the phenomenological
potential of force between sets of two quarks as
3
X p~2i X
H= + V (rij ) (9.37)
i=1
2mi i<j

r = rij = |r~1 − r~0 | (9.38)


 
3 ~2
X pi X 1 2

= + Krij + HI (9.39)

i=1
2mi i<j 2 

where H0 is the above harmonic oscillator potential and


X 1

2
HI = V (rij ) − Krij
i<j
2
X
∼ U (rij ) (9.40)
i<j

and HI is treated as the lowest order perturbation.


How do we determine HI ? In the lowest order we assume the perturbation
HI as a central term depending only upon r = rij . In QED the interaction
between two electrons is generated by mediation of a photon between the two.
It generates the Coulomb interaction in the lowest order. We assume that
QCD is just a non-Abelian interaction effect built on the lines of the Abelian
QED interaction, and hence we should have a Coulomb interaction term in
Hadron Physics 319

the one-gluon exchange interaction between two coloured quarks. Indeed one
finds that such a central term does exist in a full calculation of the one-gluon
exchange interaction (plus other terms as we shall see below). So one of the
terms in HI is 1r , a Coulomb kind of interaction.
Now one major effect of all the forces in QCD is to bring about the confine-
ment of quarks. Qualitatively we understand it this way. Let a quark-antiquark
pair be placed a distance r apart. As they are observed to be confined, it may
be assumed that a ‘string’ of strong force exists which keeps them together.
This is like, say, a magnetic dipole bound by a linear force between them.
Now when the quarks are pulled apart, the tension in the string will oppose
such an effect. We will expect the string to break if a sufficiently strong force
is applied. However as quarks are never seen singly, another q − q̄ pair would
be created between them, in Hydra-like fashion – similar to the bisection of
a magnetic dipole which at once results in two distinct magnetic dipole rods.
Hence each of these two new q − q pair is again bound. As this is what appears
to be occurring in reality, a linear potential between a confined pair of q − q̄ is
a reasonable confining force. Similarly, this should also hold for an interaction
between two quarks. Hence we assume the central HI potential should be of
the form
a
U (r) = − + br (9.41)
r
As in the case of QED, here in QCD the Coulomb interaction would leave
the first orbital excitation (N = 1) case degenerate. Thus using perturbation
theory calculations, we expect the lowest orbital baryon supermultiplets to
correspond qualitatively to the spectrum in Figure 9.2.
But even for N = 0, the nucleon N and ∆ are not degenerate. So now there
has to be more to it in U (rij) than just the central term above. Now except
for the colour factors arising due to the non-Abelian nature of QCD, the one-
gluon exchange potential in the non-relativistic limit is similar in nature to
that of the one-photon exchange in QED. Hence it is of the form,
c nc
U (rij ) = Uij (r) + Uij (r) (9.42)
where superscript ‘c’ stands for central and ‘nc’ for noncentral terms. Now,

Cαs g2
c
Uij (r) = ; αs = s (9.43)
r 4π
nc
Uij (r) = VSS (r) + VLS (r) + VT (r) (9.44)
where for various reasons VLS (r), the spin-orbit term, is found to be ignorable
in the study of hadronic spectrum and elsewhere in the quark model. And one
has therefore for the spin-spin term and the tensor terms:

ij S~i · S~j 8π 3
USS (r) = −Cαs · δ (r) (9.45)
mi mj 3
320 Group Theory in Particle, Nuclear, and Hadron Physics

(20, 1+)

(70, 2 + )

(56, 2+]

(7O, O+

(70, I" )

(56', 0+)

(56, 0+)

FIGURE 9.2: Baryon spectrum under the lowest-order perturbation theory

( )
αs 1 3(S~i · ~r)(S~j · ~r) ~ ~
UTij (r) = −C ( − Si · Sj (9.46)
mi mj r2 r2
Now for the ground state N = 0, the tensor term does not contribute and
we only need the spin-spin interaction term above.
From the previous chapter the colour factors are C = − 23 for qq in 3 of
SU (3)C baryons. This then yields C = −2 for (qqq) in colour singlet state.
For the q~q meson in colour singlet C = − 43 , these result in the lowest bound
states and other states with non-zero colours are pushed to higher energies.
Let us write Equation 9.45 as

ij S~i · S~j
VSS (r) = · Us (r) (9.47)
mi mj
where
8π 3
Us (r) = −Cαs δ (~r) (9.48)
3
(a) Mesons

~ = S~1 + S~2 So S~1 · S~2 = 1 S ~ 2 − S~1 2 − S~2 2


h i
S (9.49)
2
So,
Hadron Physics 321

ij S~i · S~j
< USS (r) >M =< > < Us (r) >M
mi mj
with Us (r) given above.

Now,
 
~ ~ 1 3
< S1 · S2 >= S (S + 1) − 2 ·
2 4
(
+ 14 f orS = 1
< S~1 · S~2 = (9.50)
− 34 f orS = 0

 
1 3 < US (r) >M < US (r) >M
m(ρ) − m(π) = − (− ) 2
= (9.51)
4 4 m m2

(b) Baryons

~ = S~1 + S~2 + S~3


S

So S = 32 , 12 , 12

!
ij S~1 · S~2 S~1 · S~3 S~2 · S~3
< USS (r) >B = + + < Us (r) >B
m1 m2 m1 m3 m2 m3
For N and ∆, all masses are equal and using,
1  ~ 2 ~ 2 ~ 2 ~ 2
S~1 · S~2 + S~1 · S~3 + S~2 · S~3 = S − S1 − S2 − S3
2
 
ij 3 < Us (r) >B
< USS (r) >B = S (S + 1) − 3 ·
4 m2
So,

3 < Us (r) >B


m(∆) − m(N ) = (9.52)
2 m2
m(∆) − m(N ) 3 < U s(r) >B 3 CB < δ 3 (r) >B
= = (9.53)
m(ρ) − m(π) 2 < U s(r) >M 2 CM < δ 3 (r) >M
3 − 2 < δ 3 (r) >B

= . 43  (9.54)
2 − 3 < δ 3 (r) >M
3
3 1 |ψ(0)|B
= . . (9.55)
2 2 |ψ(0)|3M
322 Group Theory in Particle, Nuclear, and Hadron Physics

where δ 3 (~(r)) picks just the value of meson and baryon states in the harmonic
oscillator model at the position r = 0.
Now given the ψ0S states,
 12
4α3

1 2 2
ψ0S = √ e− 2 α r
Y00 (Ω) (9.56)
π
2
|ψ0S (0)| ∼ α3 . Using α2 = b12 , where b is the scale for the harmonic oscillator,
2
and using ‘a’, the rms radius of hadrons |ψ0S (0)| ∼ a13 . So the unknown
 3
0.7 3
factor in the above is aaM

B
≈ 0.8 = 0.67. Thus (m(∆) − m(M ))/(m(ρ) −
m(π)) ∼ 0.5 and this matches well with the experiments. Thus the hyperfine
interaction is successful here.

————————————————–
Problem 9.4: Show that

(m(K ? ) − m(K)) /(m(ρ) − m(π)) = (m(Σ? ) − m(Σ)) /(m(∆) − m(N ))

m
= ms . Calculate the value of (m(Σ? ) − m(Σ)) / (m(K ? ) − m(K)).

————————————————–

The above tensor term in the one-gluon exchange potential acts only on the
S = 1 state of a pair and induces a D-state admixture in the wave function.
Overall several states get mixed with the ground state due to the above one-
gluon exchange interaction. These have been extensively studied by various
groups and yield much improved results [58].
Such models work very well in describing the hadron’s excited spectra.
It also works quite well in explaining the charge radii < r2 >n / < r2 >p
etc. Explicit calculations suggest the following successful configuration mixed
wave function.

|N >= a|56, 0+ > +b|560 , 0+ > +c|70, 0+ > d|70, 2+ > +e|20, 1+ > (9.57)

All these state are given in Table 9.2.


Its parameters are

a = 0.93, b = −0.29, c = −0.23, d = −0.04, e = 0 (in one successful set


of wave functions [70].)

and

a = 0.95, b = −0.24, c = −0.20, d = −0.042, e = −0.0024 (in another


Hadron Physics 323

successful wave function [71].)

These successful models indicate that for ∼ 80% the nucleon remains in
the N = 0 ground state configuration, while it exists for ∼ 20% in N = 2, 2hω
excited states. One normally attributes the Roper resonance N (1440) and the
J = 21 breathing-mode states of the nucleon to these N = 2 components. Note
that in these models, the D-state admixture induced in the nucleon is very
small, just about 4% or so.

9.3 Deformed Nucleon


Whether the nucleon is spherical (as assumed in the non-relativistic quark
model in Chapter 7) or is deformed, is an important topic of current studies.
The deuteron made up of a proton and a neutron has a binding energy of
2.225 MeV, J p = 1+ , I=0; magnetic moment of µd = 0.857µN , an electric
quadrupole moment of 0.256 ef m2 and a radius of 1.963 fm. So it is a very
loosely bound two-nucleon system with a large radius. A priori arguments
based on quantum mechanics predict a spherically symmetric deuteron in the
ground state.

————————————————–
Problem 9.5: Prove that in the ground state, a two-body system like a
deuteron, under the influence of a central force, should have a relative orbital
angular momentum zero.
————————————————–

However it turns out that the deuteron is not a spherically symmetric nu-
cleus. Its quadrupole moment is non-zero. Quadrupole moment operator goes
as Q ∼ r2 Y2 (θ, φ) and thus a D-state admixture (L = 2 state) is required in
the ground states. So even the simplest two-nucleon bound state is a configu-
ration mixed state.

|ψd >= a|S − state > +b|D − state >

For a2 + b2 = 1 we find an admixture of b2 = 0.04 fits the experiments.


This means that there is a 4% admixture of the D-state in the ground state of
the deuteron. Thus the deuteron is deformed in the very ground state itself.
We know that a tensor term in the two-body interaction induces the D-state
admixture.
Now the deformation of deuterons is forced upon us due to an observable
quantity – that is the non-zero quadrupole moment. And the deuteron has a
324 Group Theory in Particle, Nuclear, and Hadron Physics

non-zero quadrupole moment because its spin is 1. In contrast, the nucleon,


being a spin 12 object, has zero quadrupole moment.

————————————————–
Problem 9.6: Demonstrate that the quadrupole moment of the nucleon
or a nucleus with spin − 12 is zero.
————————————————–

Due to the above result, it was felt by many that spin-zero and spin 12
nuclei and hadrons would not be deformed. However, through high-resolution
electron scattering one obtains a tomographic view of the nucleus 164 64 Gd100 ,
which in spite of spin zero in the ground state, clearly shows a deformed
structure. This demonstrates that a zero quadrupole moment in the ground
state may still result in a deformed nucleus. To accommodate this fact all one
needs is a non-central tensor interaction to be present.
As we saw earlier the quark-quark interaction due to the one-gluon ex-
change has a tensor term and that may induce a D-state admixture in the
ground state and thereby make the nucleon deformed. However the same in-
teraction induces other admixtures too. So the admixed wave function may
very well look like that given in Equation 9.57. What is so special about the
L = 2 admixture relative to the other states like |70, 0+ >, |20, 1+ > etc.? It
is not yet clear, but one possible scenario shall be discussed later. However,
here we assume that the nucleon is indeed deformed in the ground state and
then study what advantages arise thereby.
Let us define the nucleon ground state wave function as (we are assuming
an exact SU (3)F symmetry so mu = md = ms ):
p p
|N >= 1 − PD (N )|Ns > + PD (N )|ND > (9.58)
where |NS > and |ND > correspond to the S-state and the D-state, respec-
tively, and PD (N ) is the D-state probability (for example, in the deuteron
case PD was 4% ).
Using the wave functions in Table 9.1 and Table 9.2 we have,
1
|NS >= |56, 0+ >NS = √ χρ φρ + χλ φλ ψ00S
 0
(9.59)
2
1 n  J= 12 J= 12 o
>= |70, 2+ >ND = √ φρ χS ψ2mρ 2
+ φλ χS ψ2mλ
2

|ND (9.60)
2
Using the (L-S ordered) Clebsch-Gordan Coefficient,
l X l 
X S J
|J, J3 >= χS ψ N (9.61)
m S3 J3 S3 lmΣ
m=−l S3

So here,
Hadron Physics 325

l X2 3 1

S 2 J= 1 X 3
2
[χ ψ2mX ]J3 2 = 2 2 χS2 3 ψ2mx (9.62)
m S3 J3
m=−l S3

where X = ρ or λ.
Now the spin-up and the spin-down states are
r r
 S 2 J= 12 1 S 2 1 S 2
χ ψ2mX J3 = 1 = − χ3 ψ + χ1 ψ
2 10 2 2,−1,X 5 2 2,0,X
r r
3 S 2 2 S 2
− χ 1ψ + χ 3ψ (9.63)
10 − 2 2,+1,X 5 − 2 2,2,X
And,
r r
 S 2 J= 21 2 S 2 3 S 2
χ ψ2mX J3 = −1 = − χ 3 ψ2,−2,X + χ1 ψ
2 5 2 10 2 2,−1,X
r r
1 S 2 1 S 2
− χ 1ψ + χ 3ψ (9.64)
5 − 2 2,0,X 10 − 2 2,1,X
This yields the deformed nucleon wave function in Equation 9.58.
Similar deformed ground state wave functions can be built for all the mem-
bers of the baryon octet using wave functions given in Table 9.1 and Table 9.2.
For any spin-half baryon one would use the same equations as above. Only φλ
and φρ would be different, depending upon the quark flavour structure. Take
φλ and φρ for a particular member of the spin 12 octet from Chapter 7. We
perform some calculations below.

9.4 Weak Charges


Weak charges manifest themselves in semi-leptonic decays of baryons as in
n → pe− νe− , ∧ ← pe− νe− , Σ ← pe− νe− etc. Here the ratio of the spin flip to the
GA
non-spin flip transition matrix elements is termed gA = G V
, the axial vector
coupling constant. This is because in the V-A theory of weak interactions,
the non-spin flip or Fermi transition arises from the vector (V ) interaction
and the spin-flip or Gamow-Teller transition comes from the axial vector (A)
interaction.
For the sake of completeness here, in the SU (2) spin group, the Pauli
matrices are
     
0 1 0 −i 1 0
σ1 = , σ2 = , σ3 = (9.65)
1 0 i 0 0 −1
326 Group Theory in Particle, Nuclear, and Hadron Physics

With the Lie algebra,


   
1 1 1
σi , σj = iijk σk ; (i, j, k = 1, 2, 3) (9.66)
2 2 2
As
~=
S ~
σ
2, and so
[Si, SJ ] = iyk Sk (9.67)
Define,
(σ1 ± iσ2 )
σ± = (9.68)
2
and so,
    
1 0 1 1
σ3 | ↑>= = ≡ | ↑>
0 −1 0 0
    
1 0 0 0
σ3 | ↓>= = ≡ −| ↓>
0 −1 1 −1
  
0 1 1
σ+ | ↑>= =0
0 0 0
    
0 1 1 1
σ+ | ↓>= = ≡ | ↑>
0 0 0 0
    
0 0 1 0
σ− | ↑>= = ≡ | ↓>
1 0 0 1
  
0 0 0
σ− | ↓>= =0 (9.69)
1 0 1
And the Casimir operator,
1 1
C= (σ+ σ− + σ− σ+ ) + σ32
2 4
1 2
σ + σ22 + σ32

= (9.70)
4 1
or
1
C= (S+ S− + S− S+ ) + S32 = S− S+ + S32 + S3 (9.71)
2
And

[σ+ , σ− ] = σz , [σ3 , σ+ ] = ±σ± (9.72)


Warning: These raising and lowering operators are different from the spher-
ical components σµ with µ = −1, 0, +1 in terms of normalization and signs.
1 √
σµ=±1 = ∓ √ (σ1 + iσ2 ) = ∓ 2σ± ; σ0 = σ3 (9.73)
2
Hadron Physics 327

Now SU (3)F as we saw in Chapter 3 has three SU (2) subgroups SU (2)T ,


SU (2)U , SU (2)V where the fundamental representations are, respectively,

     
u d u
; ; (9.74)
d s s
For these three SU (3) subgroups, the respective raising and lowering op-
erators are
1
τ± = (τ1 ± iτ2 ) ; τ3
2
1
u± = (u1 ± iu2 ) ; u3
2
1
v± = (v1 ± iv2 ) ; v3 (9.75)
2
where (τ1 , u1 , v1 ), (τ2 , u2 , v2 ), and (τ3 , u3 , v3 ) have similar matrices as σ1 , σ2 ,
and σ3 , respectively, as given above. The corresponding raising and lowering
operators and their effect on their respective fundamental representations are
also similar. Hence one obtains,

τ3 |u >= |u >; τ3 |d >= −|d >; τ3 |s >= 0

τ+ |u >= 0; τ+ |d >= |u >; τ+ |s >= 0

τ− |u >= |d >; τ− |d >= 0 ; τ− |s >= 0

uz |u >= 0; uz |d >= |d >; uz |s >= −|s >

u+ |u >= 0; u+ |d >= 0 ; u+ |s >= |d >

u− |u >= 0 ; u− |d >= |s > ; u− |s >= 0

vz |u >= −|u >; vz |d >= 0 ; vz |s >= |s >

v+ |u >= |s >; v+ |d >= 0 ; v+ |s >= 0

v− |u >= 0; v− |d >= 0; v− |s >= |u > (9.76)


In the weak decay of the neutron n → pe− νe−
and Σ → −
Σ0 e− νe−
at the
quark level, (udd) → (uud) and (dds) → (dus), respectively. We treat these
decays as that of the d-quark to the u-quark as τ+ |d >= |u >. The axial
vector coupling constant for a single quark, the V and the A terms are equal
and so gA for a quark is 1.
328 Group Theory in Particle, Nuclear, and Hadron Physics

 
quark GA < u ↑ |σ+ τ+ | d ↓>
gA = = =1 (9.77)
GV quark < u ↑ |τ+ | d ↑>
in SU (6)SF ⊃ SU (3)F ⊗ SU (2)S decomposition,

u↑ = u⊗ ↑, we get < u↑ |σ+ τ+ | d ↓>=< u |τ+ | d ><↑ |τ+ | ↓>= 1

Example A

For the semi-leptonic decay n → pe− ν̄e ,


P
↑ 3
< p σ (i)τ (i) n ↓>

GA i=1 + +
gA = = (9.78)
GV 3
< p↑ i=1 τ+ (i) n ↑>
P

In Chapter 7 for |NS >= |56, 0+ >N S state we obtained (the numerator
5
as 3 and the denominator as 1) :
5
gA (n → pe− ν̄e ) = (9.79)
3
This was considered to be a failure of the quark model, as this value here,
gA ∼ 0.667, is a poor fit to the experimental value of 1.25. We shall show that
a deformed nucleon fixes this perplexing problem of the quark model. Let us
calculate for the D-state,
3
X 3 1 1
< p↑D | σ+ (i)τ+ (i)|n↓D >= [< φρp |τ+ (3)|φρn >< [χS ψ2mρ
2
] 21 |σ+ (3)|[χS ψ2mρ
2 2
]− 1 >
i=1
2 2 2

1 1
+ < φλp |τ+ (3)|φλn >< [χS ψ2mλ
2
] 21 |σ+ (3)|[χS ψ2mλ
2
]−
2
1 >] (9.80)
2 2

We know already (Chapter 7),


1
< φρp |τ+ (3)| φρn >= 1 and < φλp |τ+ (3)| φλn >= − (9.81)
3
2
Using the orthonormality of the orbital state ψ2mX (x = ρ, λ), we obtain
  12  S 2  12 q q
1 3
< χS ψ2mX
2
1 |σ+ (3)| χ ψ2mX
−2 1 >= [− 10
S S
10 < χ 32 |σ+ (3)| χ 12 >
q  q  2 q q 
+ 15 − 15 < χS1 |σ+ (3)| χS− 1 > +(− 10 3
) 1 S S
10 < χ− 1 |σ+ (3)| χ− 3 >]
2 2 2 2

< χS3 |σ+ (3)| χS1 >=<↑↑↑ |σ+ (3)| √13 (↓↑↑ + ↑↓↑ + ↑↑↓) = √1
3
2 2

1 2
< χS1 |σ+ (3)| χS− 1 >= 3 <↓↑↑ + ↑↓↑ + ↑↑↓ |σ+ (3)| ↓↓↑ + ↓↑↓ + ↑↓↓>= 3
2 2

< χS− 1 |σ+ (3)| χS− 3 >= √1 <↓↓↑ + ↓↑↓ + ↑↓↓ |σ+ (3)| ↓↓↓>= √1
2 2 3 3
Hadron Physics 329

So the matrix is
q q     q  q 
1 3 √1 + √1 √1 2 3 1 √1
= − 10 10 · 3 5
− 5
· 3 + − 10 10 3
 21  S 2  12
>= − 13 (X = ρ, λ)

< χS ψ2mX
2
1 |σ+ (3)| χ ψ2mX
−1
2 2

Hence

< p↑D Σσ+ (i)τ+ (i) n↓D >= 32 1 × − 13 + − 13 − 31 = − 13
   
i

and

< p↑D |Στ+ (i)| n↑D >= 3


1 − 31 = 3 2

2 2 × 3 =1

Thus,
(1−PD ). 35 +PD (− 31 ) 5
1 − 65 PD

gA = (1−PD )+PD = 3

 5
gA n → pe− ν̄e = − 2PD (9.82)
3
With PD = 0.25 we evaluate gA = 1.26, a very good match with the ex-
periment. This is an extraordinary success of the deformed nucleon model.

Example B

For Σ− → Σ0 e− ν̄e in Chapter 7 we obtained


< Σ0↑ Σσ+ (i)τ+ (i) Σ−↓ >

i
gA = (9.83)
< Σ0↑ Στ+ (i) Σ−↑ >

i
 √ 
2 2
3 2
= √
2
= 3

This is the result for |ΣS >= |56, 0+ >ΣS .

Assuming that Σ is deformed too and its wave function is


p p
|Σ >= 1 − PD (Σ)|ΣS > + PD (Σ)|ΣD >
1 1
< Σ0↑ σ+ (i)τ+ (i)|Σ−↓ 3
φρ 0 |τ+ (3)| φρ − >< [χS ψ2mρ
2
] 21 |σ+ (3)|[χS ψ2mρ
2
]2 1 >
P
D | i D >= 2 [< Σ Σ −
2 2
1 1
+ < φλ |τ (3)|φλ
Σ0 + Σ−
>< [χS ψ2mλ
2
] 21 |σ+ (3)|[χS ψ2mλ
2
] 2 1 >]

2 2
From Chapter 7 as,
330 Group Theory in Particle, Nuclear, and Hadron Physics

< φρΣ0 |τ (3)| φρΣ− >= √12 and < φλΣ0 |τ (3)| φλΣ− >= 3√ 1
2
n o   √
2
< >D = 32 √12 − 13 + 3√ 1 1 3 √1 1
 
2
− 3 = 2 − 2×3
1 + 3 = − 3

−↑ √
< Σ0↑
S Στ+ (i) ΣS >= 2

i

Hence,
√  √ 
(1−PD ) 2 3 2 +PD − 32
− 0 −
gA (Σ → Σ e ν̄e ) = √
(1−PD ) 2+PD 2

2
gA (Σ− → Σ0 e− ν̄e ) =
− PD (9.84)
3
Here we take PD to be the same as in the n → pe− ν̄e case. In fact, it is
always the same PD in all the cases of semi-leptonic decays [75].

Example C

Λ → p e− ν̄e

< p↑ | i σ+ (i)v− (i)| Λ↓ >


P
gA = (9.85)
< p↑ | i v− (i)| Λ↑ >
P

(1 − PD ) < p↑S | i σ+ (i)v− (i)| Λ↓S > +PD < p↑D | i σ+ (i)v− (i)| Λ↓D >
P P
=
(1 − PD ) < p↑S | i v− (i)| Λ↓S > +PD < p↑D | i v− (i)| Λ↓D >
P P

< p↑S Σσ+ (i)v− (i) ∧↓S >= 32 [< χρ↑ |σ+ (3)| χρ↓ >< φρp |v− (3)| φρ∧ >

i
+ < χλ↑ |σ+ (3)| χλ↓ >< φλp |v− (3)| φλ∧ >]

Next,
ρ
< φρ
p |v− (3)| φ∧ >=
1

2
< (udu − duu) |v− (3)| − √1
2
(dsu − sdv + sud − vsd − 2uds + 2dus)
q
< √1 (udu − duu) | − √1 (0 − 0 + 0 − 0 − 2udu + 2duu) >= 2
2 2 3

and

< φλp |ν− (3)| φλΛ >=< − √16 (udu − duu − 2uud) | − 1
2 (0)) >= 0

So,
n q o q
< p↑S | ↓ 3 2 1
= 32
P 
σ
i + (i)v− (i)| Λ S >= 2 1 × 3 + − 3 × 0
q
and < p↑S | i v− (i)| Λ↑S >= 32
P
Hadron Physics 331

Next h i1 h i1
< p↑
↓ ρ
φρ χS ψ2mρ
2 2
|σ+ (3)| χS ψ2mρ
2 2
P 3
i σ+ (i)v− (i) ΛD >= 2 [< p |v− (3)| φΛ >< >

D 1 1

2 2
  12   12
+< φλp |v− (3)| φλ∧ >< χ S 2
ψ2mρ 1 |σ+ (3)| χS ψ2mρ
2
− 1 >]
2 2
nq o
= 32 2 1
= − √16

3 × − 3 × 0
q
and < p↑D | i v− (i)| Λ↑D >= 32
P

So
q
(1 − PD ) 32 + PD (− √16 )
gA = q q
(1 − PD ) 32 + PD 32

4
gA = (Λ → p e− ν̄e ) = 1 − PD (9.86)
3
Example D

Ξ− → Λe− ν̄ e

< Λ↑ | i σ+ (i)v− (i)| Ξ−↓ >


P
gA = (9.87)
< Λ↑ | i v− (i)| Ξ−↑ >
P

(1−PD )<Λ↑ −↓ ↑ P −
P
S| i σ+ (i)v− (i)|ΞS >+PD <ΛD | i σ+ (i)v− (i)|ΞD ↓>
= ↑ P
(1−PD )<ΛS | −↑ ↑ P
v− (i)|ΞS >+PD <ΛD | −↑
v− (i)|ΞD >

Now
< φρΛ |v− (3)| φρΞ− >=< − √12 (dsu − sdu + sud − usd − 2uds + 2dus) |v− (3)| −
1

2
(dss − sds) >= √16

< φλΛ |v− (3)| φΛ


Ξ− >=<
1

2
s (usd − dsu + sud − sdu − 2uds + 2dus) |v− (3)| −
1 1

6
(dss + sds − 2ssd) >= √6
So < Λ↑ | i σ+ (i)v− (i)| Ξ−↓ >= √1
P
6
q
< Λ↑S | i v− (i)| Ξ− ↑ 3
P
S >= 2

and < Λ↑D | σ+ (i)v− (i)| Ξ− ↓ √1


P
i D >= − 6
q
< Λ↑D | i v− (i)| Ξ↑D >= 32
P

whence,
1 3
gA (Ξ− → Λe− ν̄e ) = − PD (9.88)
3 2
332 Group Theory in Particle, Nuclear, and Hadron Physics
TABLE 9.3: Semi-leptonic decays of baryons with
deformation (only single parameter for all these fits)
Our Prediction
Decay Experiment Expression Value
n → pe− ν̄e 1.262±0.005 5
3 − 2PD 1.26

Λ → pe ν̄e 0.70 ±0.08 1- 43 PD 0.73
Σ− → ne− ν̄e -0.34 ±0.05 - 13 -0.33
Σ− → Λe− ν̄e (GV /GA ) 0.03±0.08 0 0
Σ− → Σ0 e− ν̄e - 2
3 − P D 0.46
Σ0 → pe− ν̄e - − 13 -0.33
0 − − 2
Σ → Σ e ν̄e - 3 − P D 0.46
Ξ0 → Σ− e− ν̄e - 5
3 − 2P D 1.26
Ξ− → Λe− ν̄e 0.248±0.06 1
3 − 2
P
3 D 0.2
Ξ0 → Σ0 e− ν̄e - 5
3 − 2P D 1.26
Ξ− → Ξ0 e− ν̄e - − 13 -0.33

————————————————–
Problem 9.7: Show that
(a) gA (Σ− → ne− ν̄e )) = − 13

(b) gA (Ξ0 → Σ+ e− ν̄e ) = 5


3 − 2PD

(c) gA (Ξ− → Σ0 e− ν̄e ) = 5


3 − 2PD

(d) gA (Ξ− → Ξ0 e− ν̄e ) = − 13

(e) gA (Σ0 → pe− ν̄e ) = − 31

(f) gA (Σ0 → Σ+ e− ν̄e ) = 2


3 − PD
GV −
(g) GA (Σ → Λe− ν̄e ) = 0

(Note relevant changes in the last ratio.)


————————————————–

We give our results for the semi-leptonic decay of baryons in Table 9.3.
We utilize one single parameter PD = 0.208 to predict all the values in our
model. Agreement with experimental measurements is good. This supports
the argument that all the spin − 12 octet baryons are deformed in the ground
state itself.
Hadron Physics 333

9.5 Magnetic Moments


We define the magnetic moment operator in the quark model as
3
X
µ= (Q(i)σz (i) + Q(i)lz (i)) µ0 (9.89)
i=1

where
e~
µ0 = (9.90)
2mq c
Here we have used the SU (3) symmetry for the same mass u- and d- quarks
mq = mu = md . We have utilized the above expression without the lz term to
calculate the magnetic moment of baryons in Chapter 7 for the spherical quark
model for | 56, 0+ > states. Now we extend the same to include a deformed
nucleon or baryon as we defined above. The new term to be included is lz .
From our definition of the Jacobi coordinates of the three-quark system note
that,

l~ρ = ρ
~ × p~ρ ; l~λ = ~λ × p~λ (9.91)
With respect to the centre of mass, the coordinate of position 3 is

r
~ 1 2~
R3 = r~3 − (r~1 + r~2 + r~3 ) = − λ
3 3
So,
2~
Lz = L3 = R~3 × P~3 = Lλ (9.92)
3
2
L~1 + L~2 = ~lλ + ~lρ (9.93)
3
L = L1 + L2 + L~3 = ~lρ + ~lλ
~ ~ ~ (9.94)
Hence for the neutron spin-up state; taking first the σz term,
3
< N ↑ |µ|N ↑ >= (1 − PD (N )) < NS↑ | σz (i)Q(i)|NS↑ >
X

i=1
3
↑ ↑
X
+PD (N ) < ND | σz (i)Q(i)|ND > (9.95)
i=1

3
< NS↑ |Σσz (i)Q(i)|NS↑ > [< φρn |Q(3)|φρn >< χρ↑ |σz (3)|χρ↑ >
2
334 Group Theory in Particle, Nuclear, and Hadron Physics

+ < φλn |Q(3)| φλn >< χλ↑ |σz (3)| χλ↑ >]
Given,
1 1
< φρn |Q(3)| φρn >= − ; < φλn |Q(3)| φλn >=
3 3
1
< χρ↑ |σz (3)| χρ↑ >= 1 < χλ↑ |σz (3)| χλ↑ >= − (9.96)
3
2
< NS↑ |Σσz (i)Q(i)| NS↑ >= −
3
X 3 h i1 h i1
↑ ↑ 2 2
< ND | σz (i)Q(i)|ND >= [< φρn |Q(3)| φρn >< χS ψ2mρ
2
|σz (3)| χ S 2
ψ 2mρ >
i
2 1
2
1
2

 12  S 2  12
+ < φλn |Q(3)| φλn >< χS ψ2mλ
2

1 |σz (3)| χ ψ2mλ 1 >]
2 2

Given that,
 21  S 2  12 1
< χS ψ2mX
2

1 |σz (3)| χ ψ2mX 1 >= − ; (X = ρ, λ) (9.97)
2 2 3
↑ ↑
< ND |Σσz (i)Q(i)| ND >= 0 (9.98)
Thus,
2
< N ↑ |Σσz (i)Q(i)| N ↑ >= − (1 − PD (N )) (9.99)
3
Next similar term for proton:
X
< P↑ σz (i)Q(i) P ↑ >

X X
= (1 − PD (N )) < PS↑ σz (i)Q(i) PS↑ > +PD (N ) < PD
↑ ↑
σz (i)Q(i) PD >)


(9.100)
Given,
2
< φρp |Q(3)| φρp >= ; < φλp |Q(3)| φλp >= 0 (9.101)
3
We obtain,

< PS↑ Σσz (i)Q(i) PS↑ >= 1 ↑ ↑
< PD Σσz (i)Q(i) PD >= 0

i i

So, X 4 N
< P↑ σz (i)Q(i) P ↑ >= 1 − PD (9.102)

3
If we were to ignore the lz term above then,

3 1 − 43 PD (N )

µP (σ3 Q)
=−
µN (σ3 Q) 2 (1 − PD (N ))
Hadron Physics 335

This is a bad fit, as it takes us away from the earlier good fit of the spherical
quark model result of − 32 . However this only shows that one cannot ignore
the lz term. Hence let us include this term.
Now,
1
| 56, 0+ >N S = √ φρ χρ + φλ χλ ψ00S
 0
2
1
 4α3 2 − 1 α2 r2 0

1 ρ ρ λ λ
= √ φ χ +φ χ √ e 2 Y0 (Ω) (9.103)
2 π
Now lz acting as Y00 (Ω) gives zero and hence for both the proton and the
neutron,

lz | 56, 0+ >N S = 0 (9.104)


Hence no contribution for the lz term arises from the S-state, and hence
only the D-state will contribute. Note lz = 32 Lλ .
Hence we have
r
8 λ5  2 2 2 2 2
2
ρ Ym (Ωρ ) − λ2 Ym2 (Ωλ ) e−α (ρ +λ )/2

ψ2mλ =
15 π
Now the action of lz on it is
r
8 λ5  2 2 2 2
2
ρ lλ Ym2 (Ωρ ) − λ2 lλ Ym2 (Ωλ ) e−α (ρ +λ )/2

lλ ψ2mλ =
15 π
r
8 λ5  2 2 2
0 − mλ2 Ym2 (Ωλ ) e−α (ρ +λ )/2

=
15 π
So,
2 2 m
< ψ2mλ |lλ | ψ2mλ >= (9.105)
2
Therefore, using Table 9.2 we obtain
h i1 h i1 1 1
2 2
< χS ψ2mλ
2
|lλ | χS ψ2mλ
2
>= 2
< ψ2−1λ 2
|lλ | ψ2−1λ 2
> + < ψ20λ 2
|lλ | ψ20λ >
1
2
1
2
10 5
3 2 2 2 2 2
+ < ψ21λ |lλ | ψ21λ > + < ψ22λ |lλ | ψ22λ >
10 5
 
1 1 1 3 1 2 2
= × − + ×0+ × + ×
10 2 5 10 2 5 2
 21  S 2  12 1
< χS ψ2mλ
2

1 |lχ | χ ψ2mλ 1 >= (9.106)
2 2 2
Next,
8 λ5 X 1 
1 2 2 2 2
2
ψ2mρ = ρλ Yµ1 (Ωρ )Yµ10 (Ωλ )e−α (ρ +λ )/2 (9.107)
3 π 21 0
µ µ0 m
µµ
336 Group Theory in Particle, Nuclear, and Hadron Physics

So,
2 8 α5 1 1 −α2 (ρ2 +λ2 )/2
ψ22ρ = 1 ρλY1 (Ωρ )Y1 (Ωλ )e
3 π2
lλ acting on it gives +1 and so,
2 2 2 2
< ψ22ρ |lλ | ψ22ρ >= +1, < ψ2−2ρ |lλ | ψ2−2ρ >= −1

8 α5
 
2 1 1 2 1
ψ21ρ = ρλ[ Y (Ω )Y 1 (Ω )
3 π 21 1 0 1 1 ρ 0 λ
 
1 1 2 1 2 2 2
+ Y (Ω )Y 1 (Ω )e−α (ρ +λ )/2 ]
0 1 1 0 ρ 1 λ
Now,

8 α5
 
2 1 1 1 −α2 (ρ2 +λ2 )/2
lλ ψ21ρ = ρλ 0 + √ Y0 (Ω )Y
ρ 1 (Ω λ )e
3 π 21 2

2 2 1 2 2 1 2 2
< ψ21ρ |lλ | ψ21ρ >= , < ψ2−1ρ |lλ | ψ2−1ρ >= − , < ψ20ρ |lλ | ψ20ρ >= 0
2 2
and so,
 12  S 2  12 1
< χS ψ2mρ
2
< χ22−1ρ |lλ | χ22−1ρ >

1 |lλ | χ ψ2mρ 1 >=
2 2 10

1 2 2 3 2 2 2 2 2
+ < ψ20ρ |lλ | ψ20ρ >+ < ψ21ρ |lλ | ψ21ρ > + < ψ22ρ |lλ | ψ22ρ >
5 10 5

 12  S 2  21 1
< χS ψ2mρ2

1 |lλ | χ ψ2mρ 1 >= (9.108)
2 2 2
Next we calculate the ~l contribution to the magnetic moment. First for the
neutron,

X


↑ 2 ↑ ↑
< ND Q(i)lz (i) ND >= 3 · < ND |Q(3)lλ | ND >

i
3

2 1  12  S 2  12
= 3. · [< φρn |Q(3)| φρn >< χS ψ2mρ
2

1 |lλ | χ ψ2mρ 1 >
3 2 2 2

 12  S 2  12
+ < φλn |Q(3)| φλn >< χS ψ2mλ
2

1 |lλ | χ ψ2mλ 1 >]
2 2

  
2 1 1 1 1 1
=3· · − × + × =0
3 2 3 2 3 2
Hadron Physics 337

Thus for neutrons one finally obtains



↑ ↑
< ND ΣQ(i)lz (i) ND >= 0 (9.109)
i

Next for the proton spin up D-state,




↑  12  S 2  21
ΣQ(i)lz (i) PD >∼ [< φρp |Q(3)| φρp >< χS ψ2mρ
2

< PD 1 |lλ | χ ψ2mρ 1 >
i 2 2

 12  S 2  12 2 1 2 1 1
+ < φλp |Q(3)| φλp >< χS ψ2mλ
2

1 |lλ | χ ψ2mλ 1 >] = 3. . ( · + 0) =
2 2 3 2 3 2 3
Therefore,
1
< P ↑ ΣQ(i)lz P ↑ >= PD (N ) (9.110)

i 3
So the neutron has no contribution from ~l terms while the proton does
have PD (N )/3 from the l-term. So finally adding up spin and ~l terms,
2
µN = − (1 − PD (N )) µq (9.111)
3
µp = (1 − PD (N )) µq (9.112)
Whence,
µN 3
=− (9.113)
µp 2
This is the same result as for the spherical quark model. However here the
individual terms have a (1 − PD (N )) factor in each case, which cancels out in
the ratio above. Consequently the deformed baryon maintains the good fits of
µN
µp , but now with extra term PD (N ) which goes out in the ratio but remains
as a renormalization term (for mass mq ) as we see below.
One more example: the magnetic moment of Ξ0 is

< Ξ0 |µZ | Ξ0 >= (1 − PD ) < Ξ0S |µZ | Ξ0S > +PD < Ξ0D |µZ | Ξ0D > (9.114)

We already know for the spin term µS = σz Q,


X 4 4 e~
< Ξ0S | σz (i)Q(i)µ0 (i)|Ξ0S >= (− − )
i
9ms 9mu 2c
X
< Ξ0D σ(i)Q(i)µ0 (i) Ξ0D >


3 Q(3) ρ  12  S 2  12
[< φρΞ0 φ 0 >< χS ψ2mρ
2

= Ξ 1 |σ3 (3)| χ ψ2mρ 1 >
2 mq (3) 2 2
338 Group Theory in Particle, Nuclear, and Hadron Physics


Q(3) ρ  S 2  12  S 2  12 e~
+< φλΞ0 mq (3) φΞ0 >< χ ψ2mλ |σz (3)| χ ψ2mρ >] 2c
(9.115)

 
2 1 1 e~
=− − (9.116)
9 mu ms 2c
Next,
< Ξ0D |µz (l3 )| Ξ0D >

3 ρ Q(3) ρ  12  S 2  12
φΞ0 >< χS ψ2mρ2

= [< φΞ0 1 |lz (3)| χ ψ2mρ 1 >
2 mq (3) 2 2


Q(3) λ  12  12 e~
+ < φλΞ0 φ 0 >< χS ψ2mλ2
|lz (3)| χS ψ2mλ
2
 
Ξ 1 >]
mq (3) 2 2c
 
2 1 1 e~
= −
9 mu ms 2c
So,

< Ξ0D |µ|Ξ0D >=< Ξ0D |µz (σ)|Ξ0D > + < Ξ0D |µz (l)|Ξ0D >
   
2 1 1 2 1 1
=− − + − =0 (9.117)
9 mu ms 9 mu ms
Finally,
 
2 4 e~
< Ξ0 |µz | Ξ0 >= (1 − PD ) − − (9.118)
9mu 9ms 2c

Next we show that the sum of the spin plus orbital component for a de-
formed spin 21 baryon is generic in the quark model. In fact we see here that,

3

Q(3) ρ h i1 h i1
2 2
< Ξ0D |µz (σ)| Ξ0D >= [< φρΞ0 φ 0 >< χS ψ2mρ
Ξ
2
|σz (3)| χS ψ2mρ
2
>
2 mq (3) 1
2
1
2

Q(3) λ  12  12 e~
+ < φλΞ0 φ 0 >< χS ψ2mλ
2
|σz (3)| χS ψ2mλ
2
 
Ξ 1 >]
mq (3) 2 2c


3 1 Q(3) ρ
φ 0 + < φλ 0 Q(3) φλ 0 >] e~
(− )[< φρΞ0

= Ξ (9.119)
2 3 mq (3) Ξ mq (3) Ξ 2c

While for
   
3 2 1 e~
< Ξ0D |µz (l)| Ξ0D >= · · {< | | >ρ + < | | >χ } (9.120)
2 3 2 2c
Hadron Physics 339

The spin and the orbital components contribute the same amount with
opposite signs and thus the sum is zero. Now this holds for any member of
the spin 12 octet baryons. Hence the result is generic.

< Ξ0D |µ(”σ + l”)| Ξ0D >= 0 (9.121)

————————————————–
Problem 9.8: Demonstrate that for the deformed Σ+ the magnetic mo-
ment is  
8 1 e~
< Σ+ |µ| Σ+ >= (1 − PD ) + (9.122)
9mu 9ms 2c
————————————————–
Unsolved Problem 9.2: Obtain the magnetic moments of deformed
baryons Λ, Σ0 , Σ− and Ξ− . Confirm that these differ from the spherical baryon
case only by a multiplicative factor (1 − PD ).
————————————————–

Now note that we find that generically for the spin 21 deformed states of
the octet, the magnetic moment ratio is the same as that of the corresponding
spherical space. But all these magnetic moments we normalize by the same
factor (1 − PD ). What does it mean? We suggest that this means that the
quark masses are normalized as
mq
m= (9.123)
1 − PD
where the mass m is the renormalized mass. For PD = 0 m = mu = md ∼
340M eV , the constituent quark mass of the non-relativistic spherical quark
mass. For PD ∼ 41 (the single value of the deformation parameter which fits a
large number of physical quantities) mq = 255 MeV. These provide a success
of the deformed nucleon model in several studies, such as the M1 transition
moment for ∆+ → pγ [76], and the double delta coupling required in (π, 2π)
reactions [77],[78].

9.6 Spin of Nucleon in the Quark Model


In the quark model there are only valence quarks and no antiquarks and no
s quarks and hence the expressions for the integrated spin dependent structure
functions become

Z 1  
1 4 ↑  1 ↑
g1n = u (x) − u↓ (x) + d (x) − d↓ (x)

dx (9.124)
2 0 9 9
340 Group Theory in Particle, Nuclear, and Hadron Physics

1 1
Z  
4 ↑  1 ↑
g1n = d (x) − d↓ (x) + u (x) − u↓ (x)

dx (9.125)
2 0 9 9
R1
where u↑ = 0 u↑ (x)dx, etc. Note that we have used isospin invariance where
the u-population in the proton is equal to the d-population in the neutron. So
u↑ is that for proton in both the expressions g1 part g1n .
We consider two quark model wave functions which have been successful
in describing the ground state properties of the nucleon, however which are of
unpolarized nature. The two models that we discuss here are: (1) the deformed
nucleon model which we discussed in detail in a previous chapter and which
has unequivocally been successful in describing a large number of observables,
and next (2) the configuration mixed quark models which have been very
successful in describing the spectroscopic properties of hadrons.
But as we discussed earlier, success in describing unpolarized physical ob-
servables also ensures success in fitting the fully integrated structure functions
of an unpolarized nature. Those are pretty liberal in allowing for a large num-
ber of quark models (including the relativistic ones ) to explain the unpolarized
experimental data reasonably well. However in contrast, the polarized physical
observables are more demanding and much more discriminating, as we shall
see below.

9.6.1 Spin of a Deformed Nucleon


As we saw earlier, the addition of a D-state admixture to the spherical
states of the nucleon in the ground state itself leads to major improvements
in fitting the quark model to a large number of observables. Herein we study
this model to understand the spin structure of the nucleon.
Define the nucleon wave function as:
p p
| N >= 1 − PD | N > + PD | N > (9.126)
where | NS > and | ND > are the (56, 0+ ) and (70, 2+ ) parts of the SU (6)SF
wave function and PD is the D-states probabilities of the nucleon.
Now,
1
| NS >= √ (φρ χρ + φλ χλ ) (9.127)
2
where these states of (56, 0+ ).
Now to obtain u↑ ,u↓ d↑ and d↓ , we define the operator,
X
ÔF,S3 = PF (i)PS3 (i) (9.128)
i=1,2,3

where i labels the position of the quarks in the wave function, F stands for
flavour of u and d quarks and S3 spin ↑or ↓ of the quarks, both at the ith
position. So Pu (3) acting on quarks will pick up only u quarks if it is located
at position 3.
Hadron Physics 341

Now,
Pu (3) | u(3) >= + | u(3) > and Pu (3) | d(3) >= 0
P↑ (3) | q ↑ (3) = + | q ↑ (3)
P↑ (3) | q ↓ (3) = 0 (9.129)
So by symmetry for example,
X
u↑ =< p↑ | ΣPu (i)P↑ (i) | p↑
i=1,2,3

= 3 < p↑ | Pu (3)P↑ (3)p↑ > (9.130)


R0
To extract u↑ = 0 u↑ (x)dx we can obtain it in different ways, depend-
ing upon how we define the appropriate operator. First to obtain values for
|nS > (56, 0+ ) ground states, let us define an operator Nu↑ , the number of
u-quarks with spin up in the wave function. So |nS >=| pS >, a proton in
S-state:

u↑p =< ps↑ |Nu↑ | ps↑ >


1
= [< φpρ χ↑ρ | Nu↑ | φpρ χ↑ρ > + < φpλ χ↑λ | Nu↑ | φpλ χ↑λ >
2
+ < φpρ χ↑ρ | Nu↑ | φλρ χ↑λ > + < φpλ χ↑λ | Nu↑ | φλρ χ↑ρ >] (9.131)
Note there is nothing from the last two cross terms to contribute. (Note
1
−1
spin up ↑= χ 21 and ↓= χ 1 2 for spin down states),
2 2

< φpρ χ↑ρ | Nu↑ | φpρ χ↑ρ >

1 1
=< (udu − duu) (↑↓↑ − ↓↑↑) |Nu↑ | (udu − duu) (↑↓↑ − ↓↑↑) >
2 2
Note

nu↑ | udu (↑↓↑ − ↓↑↑) − duu (↑↓↑ − ↓↑↑) >

= nu↑ | u↑ d↓ u↑ − u↓ d↑ u↑ − d↑ u↓ u↑ − d↓ u↑ u↑ >

= 2u↑ d↓ u↑ − u↑ d↑ u↓ − d↑ u↓ u↑ − 2d↓ u↑ u↑ >

= udu (2 ↑↓↑ − ↓↑↑) − duu (↑↓↑ −2 ↓↑↑)


(Note first we express (udu) (↑↓↑) as u↑ d↓ u↑ . If in doubt start by placing


position labels in wave functions. These are


342 Group Theory in Particle, Nuclear, and Hadron Physics

{u(1)d(2)u(3)} {↑ (1) ↓ (2) ↑ (3)}


And thereafter we link flavour and spin residing at the same position as
u(1) ↑ (1) = u↑ (1)).
So
1
< φpρ χ↑ρ | Nu↑ | φpρ χ↑ρ >= [< udu (↑↓↑ − ↓↑↑) − duu (↑↓↑ − ↓↑↑)
4

| udu (2 ↑↓↑ − ↓↑↑) − duu (↑↓↑ −2 ↓↑↑) >]

1 3
= (2 + 1 + 1 + 2) =
4 2
And
1
< φpλ χ↑λ | Nu↑ | φpλ χ↑λ >=< (udu + duu + 2uud) (↑↓↑ + ↓↑↑ −2 ↑↑↓)
6
1
| [udu (2 ↑↓↑ + ↓↑↑ −2 ↑↑↓) + duu (↑↓↑ +2 ↓↑↑ −2 ↑↑↓) − 2uud (↑↓↑ + ↓↑↑ −4 ↑↑↓)]
6
Note that the last term is 2 × 2 ↑↑↓ as it is with uud in flavour space and we
have two u-quarks with spin-up in this case. Thus we are counting spin-up
terms in the spin part while leaving flavour terms free. We may also count in
the flavour space and leave the spin space free. Thus,
3
=
2
Now the two cross terms should be equal to each other and

1
< φpλ χ↑λ | Nu↑ | φpλ χ↑λ >=< (udu + duu − 2uud) (↑↓↑ + ↓↑↑ −2 ↑↑↓) >
6

1 1
|
{udu (2 ↑↓↑ + ↓↑↑) − duu (↑↓↑ +2 ↓↑↑)} =
2 6
Putting these together above,
5
u↑p =< p↑S | Nu↑ | p↑S >=
3
Form u ↔ d exchange symmetry between protons and neutrons,
5
d↑n =< nS ↑ | Nd↑ | nS ↑ >=
3
Next we obtain u↑ in the neutron,

u↑n =< ns↑ | Nu↑ | ns↑ >


Hadron Physics 343

1
= [< φnρ χ↑ρ | Nu↑ | φnρ χ↑ρ > + < φnλ χ↑λ | Nu↑ | φnλ χ↑λ >
2

+ < φnρ χ↑ρ | Nu↑ | φnλ χ↑λ > + < φnλ χ↑λ | Nu↑ | φnρ χ↑ρ >]
Now,
1
< φnρ χ↑ρ | Nu↑ | φnρ χ↑ρ >= (udd − dud) (↑↓↑ + ↓↑↑)
2

1
| Nu↑ | (udd − duu) (↑↓↑ + ↓↑↑) >
2

1 1 1
=< (udd − duu) (↑↓↑ + ↓↑↑) | (udd(↑↓↑ −0) − dud(0− ↓↑↑) >=
2 2 2
Next
1
< φnλ χ↑λ | Nu↑ | φnλ χ↑λ >=< − (udd + dud − 2ddu) (↑↓↑ + ↓↑↑ −2 ↑↑↓)
6

1
[− udd (↑↓↑ +0 − 2 ↑↑↓) + dud (0+ ↓↑↑ −2 ↑↑↓) − 2ddu (↑↓↑ + ↓↑↑ −0)]
6

1
=
2
and
1
< φnλ χ↑λ | Nu↑ | φnρ χ↑ρ >< − (udd + dud − 2ddu) (↑↓↑ + ↓↑↑ −2 ↑↑↓) >
6

1 1
| [ (udd) (↑↓↑) + (dud) (↓↑↑)] >= −
2 6
1
u↑n =< nS ↑ | Nu↑ | nS ↑ >= (9.132)
3
By symmetry,
1
d↑p =< pS ↑ | Nd↑ | pS ↑ >= (9.133)
3

————————————————–
Unsolved Problem 9.3: Prove the following
1
u↓p = d↓n = (9.134)
3
344 Group Theory in Particle, Nuclear, and Hadron Physics

2
d↓p = u↓n = (9.135)
3
————————————————–

Note for consistency we need


5 1
u↑p + u↓p = + =2 (9.136)
3 3

2 1
d↑p + d↓p = + =1 (9.137)
3 3
This is good as there are two-u and one-d quark in the proton.
So finally,
 
p 1 4 ↑ ↓
 1 ↑ ↓

g1 = u −u + d −d (9.138)
2 9 9
    
1 4 5 1 1 1 2
= − + −
2 9 3 3 9 3 3
5
g1p = (9.139)
18
 
1 4 ↑  1 ↑
g1n = d − d↓ + u − u↓

(9.140)
2 9 9
    
1 4 1 2 1 5 1
= − + −
2 9 3 3 9 3 3

g1n = 0 (9.141)
Note that the Bjorken sum rule is
Z 0
1
(g1p − g1n ) dx = (gA /gV ) (9.142)
0 6
Using the above values, this sum rule gives gA /gV = 53 , the non-relativistic
quark model result for quarks in the S-state, (56, 0+ ) representation.
Next let us include the D-state for the proton,

1 n   21  s 2  21 o
| PD >≡| 70, 2+ >p = √ φρp χs ψ2mρ
2 λ
1 + φp χ ψ2mλ 1 (9.143)
2 2 2

2 2
Note that as < ψ2mρ | ψ2mλ >= 0, there will be no cross terms in the
matrix element:
Hadron Physics 345

1  12  S 2  12
< p↑D | Nu↑ | p↑D >= [< φρp χS ψ2mρ
2 ρ

1 |Nu↑ | φp χ ψ2mρ 1 >
2 2 2

 12  S 2  12
+ < φλp χS ψ2mλ
2 λ

1 |Nu↑ | φp χ ψ2mλ 1 >]
2 2

Now,
 12  S 2  12
< φρp χS ψ2mρ
2 ρ

1 |Nu↑ | φp χ ψ2mρ 1 >
2 2

1 1
= < φρp χS3 |N u↑ | φρp χS3 > + < φρp χS1 |N u↑ | φρp χS1 >
10 2 2 5 2 2

3 2
+ < φρp χS− 1 |N u↑ | φρp χS− 1 > + < φρp χS− 3 |N u↑ | φρp χS− 3 >
10 2 2 5 2 2

Where
< φρp χS3 |N u↑ | φρp χS3 >
2 2

1 1
=< √ (udu − duu) ↑↑↑ |N u↑ | √ (udu − duu) ↑↑↑>= 2
2 2

1 1
< φρp χS1 |N u↑ | φρp χS1 >=< √ (udu − duu) √ (↑↑↓ + ↑↓↑ + ↓↑↑) |
2 2 2 3

 
1 4
√ udu (↑↑↓ +2 ↑↓↑ + ↓↑↑) − duu (↑↑↓ + ↑↓↑ +2 ↓↑↑) =
6 3

1 1
< φρp χS− 1 |N u↑ | φρp χS− 1 >=< √ (udu − duu) √ (↓↓↑ + ↓↑↓ + ↑↓↓) |
2 2 2 3
 
1 2
√ udu (↓↓↑ +0+ ↑↓↓) − duu (↓↓↑ + ↓↑↓ +0) =
6 3

< φρp χS− 3 |N u↑ | φρp χS− 3 >= 0


2 2

Therefore,
 21  S 2  12 2
< φρp χS ψ2mρ
2 ρ

1 |Nu↑ | φp χ ψ2mρ 1 >=
2 2 3
Next,
 12  s 2  12
< φλp χs ψ2mλ
2 λ

1 |Nu↑ | φp χ ψ2mλ 1 >
2 2

1 1
= < φλp χS3 |N u↑ | φλp χS3 > + < φλp χS1 |N u↑ | φλp χS1 >
10 2 2 5 2 2
346 Group Theory in Particle, Nuclear, and Hadron Physics

3 2
+ < φλp χS− 1 |N u↑ | φλp χS− 1 > + < φλp χS− 3 |N u↑ | φλp χS− 3 >
10 2 2 5 2 2

We find,
4
< φλp χS3 |N u↑ | φλp χS3 >= 2; < φλp χS1 |N u↑ | φλp χS1 >=
2 2 2 2 3

2
< φλp χS− 1 |N u↑ | φλp χS− 1 >= ; < φλp χS− 3 |N u↑ | φλp χS− 3 >= 0
2 2 3 2 2

Therefore,
2
< p↑D | Nu↑ | p↑D >=
3
Thus,

u↑p =< p↑ | Nu↑ | p↑ >=< (1 − PD ) < p↑S | Nu↑ | p↑S > +PD < p↑D | Nu↑ | p↑D >

5
u↑p = − pD (9.144)
3

1  12  S 2  12
d↑p =< p↑D | Nd↑ | p↑D >= [< φρp χS ψ2mρ
2 ρ

1 |Nd↑ | φp χ ψ2mρ 1 >
2 2 2

 12  S 2  12
+ < φλp χS ψ2mλ
2 λ

1 |Nd↑ | φp χ ψ2mλ 1 >]
2 2

Now
 12  S 2  12
< φρp χS ψ2mρ
2 ρ

1 |Nd↑ | φp χ ψ2mρ 1 >
2 2

1 1
= < φρp χS3 |N d↑ | φρp χS3 > + < φρp χS1 |N d↑ | φρp χS1 >
10 2 2 5 2 2

3 2
+ < φλp χS− 1 |N d↑ | φλp χS− 1 > + < φλp χS− 3 |N d↑ | φλp χS− 3 >
10 2 2 5 2 2

So,
1
< p↑D | Nd↑ | p↑D >=
3
And,

d↑p =< p↑ | Nd↑ | p↑ >=< (1 − PD ) < p↑S | Nd↑ | p↑S > +PD < p↑D | Nd↑ | p↑D >

1 1
= (1 − PD ) × + PD ×
3 3
Hadron Physics 347

Finally,
1
d↑p = (9.145)
3

————————————————–
Problem 9.9: For a deformed neutron show that
1 5
and d↑n = − PD
u↑n = (9.146)
3 3
————————————————–
Unsolved Problem 9.4: Using the above technique for the deformed
nucleon demonstrate that
4 2
, < p↑D | Nd↓ | p↑D >=
< p↑D | Nu↓ | p↑D >=
3 3
————————————————–

And so, Z 1
1
u↓p = u↓ (x)dx = + PD (9.147)
0 3
Z 0
2
d↓p = d↓ (x)dx = (9.148)
0 3
Putting all the relevant quantities together,
 
p 1 4 ↑ ↓
 1 ↑ ↓

g1 = u −u + d −d
2 9 9
    
1 4 5 1 1 1 2
= − pD − − pD + −
2 9 3 3 9 3 3
5 4
g1p =
− PD (9.149)
18 9
    
1 4 1 2 1 5 1
g1n = − + − PD − − PD
2 9 3 3 9 3 3
1
g1n = − PD (9.150)
9
Note that as it should, the Bjorken sum rule is satisfied,
 
1 gA
g1p − g1n =
6 gV
PD = 0.203 gives g1p = 0.187 and g1n = −0.023
348 Group Theory in Particle, Nuclear, and Hadron Physics

PD = 0.25 gives g1p = 0.167 and g1n = −0.028

PD = 0.3 gives g1p = 0.144 and g1n = −0.033

PD = 0.33 gives g1p = 0.129 and g1n = −0.037

where the last one matches the experimental value [57] of g1p = 0.129 and g1n =
−0.031 more closely. Thus a suitably deformed nucleon gives reasonable fits
to the experimental numbers and indicates that the resolution of the proton
spin crises is to be found in the large contribution from the orbital angular
momentum in the quark model.
Now this quark model is not of the constituent quark kind which yields
g1p = 18 5
= 0.278 and g1n = 0, both of which fail to match experimental
quantities. g1p = 18
5
= 0.278 is pretty bad as it is. But g1n = 0 is a complete
and utter disaster, as genetically due to the goodness of the Bjorken sum
rule, it is guaranteed to be zero in the constituent quark model. Only a quasi-
quark model of the deformed nucleon kind matches these polarized physical
observables.

9.6.2 Spin of Nucleon with Configuration Mixed Wave Func-


tion
Good fits to baryon spectra and decay properties are obtained by configu-
ration mixed wave functions in the quark model. The mixing is induced in the
harmonic oscillator model by an an-harmonic term and a hyperfine interac-
tion. This was considered a major success of these configuration mixed wave
functions [71],[70] within the SU (6)SF ⊗ SO(3) symmetry group.
This hyperfine function interaction also induce N = 2 configuration mixing
in the N = 0 ground state wave function. Explicit calculations suggest the
following wave functions

| N >= a | 56, 0+ > +b | 560 , 0+ > +c | 70, 0+ >


+d | 70, 2+ > +e | 20, 1+ > (9.151)
where a = 0.93, b = −0.29, c = −0.23, d = −0.04, e = 0 in [70],

and,

a = 0.95, b = −0.24, c = −0.20, d = −0.042, e = −0.0024 in [71].

These models indicate that of about ∼ 80% of the time the nucleon remains
in the ground state configuration, while it exists only for 20% of the time in
the N = 2, 2~ω excitations.
Now the tensor term in the above hyperfine interaction may induce D-state
admixture, as for the deformed nucleon case discussed above. But this tensor
Hadron Physics 349

term between quarks may, at best, induce only a few percent of the D-state.
This is therefore not enough to explain as large a deformation as PD ∼ 0.25.
Thus the quark-quark tensor interaction here is not strong enough to deform
the nucleon as strongly as we require it to fix the physical quantities as we
discussed above. Therefore it is clear that some other mechanism is required
to accomplish this task. What it is, we discuss below.
Having shown how the deformed nucleon successfully explains the polar-
ized structure functions of proton and neutron, let us see how these successful
(for the unpolarized physical observables) configuration mixed wave functions
fare as to the same polarized physical observables.
We have worked out u↑ , u↓ , d↑ and d↓ for the N = 0 state | 56, 0+ >. Using
N
notation of Table 9.2 for ψlmΣ (Σ = S, A, ρ, λ) for the ground state,

α3 2 2 2
0
ψ00S = 3 e−α (ρ +λ )/2 (9.152)
π 2

Now for the state | 560 , 0+ >, which is an N = 2 state,

1 α3  2 2 2
2
= √ 3 3 − α2 ρ2 + λ2 e−α (ρ +λ )/2

ψ00S (9.153)
3 π2
Here
1 1
| 560 , 0+ >J= 2 = √ χρ φρ + χλ φλ ψ00s
 0
(9.154)
2

————————————————–
0 2
Problem 9.10: Verify the normalization of the state ψ00S and ψ00S .
————————————————–

0 2
There should not be any cross terms < ψ00S | ψ00S > as these states are
orthonormal. Check
ZZ
2? 2
ψ00s ψ00s d3 ρd3 λ

1 α3 α3
   ZZ  
2 2 2
3 − α2 ρ2 + λ2 e−α (ρ +λ ) ρ2 λ2 dρdλ
2 
= √ 3 3 (4π)
3 π2 π 2

Z ∞ Z ∞
2 2 2
3ρ2 λ2 − α2 ρ4 λ2 − α2 ρ2 λ4 e−α (ρ +λ ) dρdλ

= (..)
0 0

Z ∞ Z ∞ Z ∞ Z ∞
2 2 2
λ2 2 2 2
λ2
= (.)[3 ρ2 e−α ρ
dρ λ2 e−α dλ − α2 ρ4 e−α ρ
dρ λ2 e−α dλ
0 0 0 0
350 Group Theory in Particle, Nuclear, and Hadron Physics

Z ∞ Z ∞
2 2 2
λ2
−α2 ρ2 e−α ρ
dρ λ4 e−α dλ]
0 0

√ √ √ √ √ √
1 π 1 π 3 π1 π 21 π 3 π
[.] = 3 . 3 . . 3 . − α2 − α . =0 (9.155)
4 α 4 α 8 α5 4 α3 4 α3 8 α5
Hence the state | 56, 0+ > has the same results as that of | 56, 0+ >.
5 1 1 2
u↑ = , u↓ = , d ↑ = , u↓ = (9.156)
3 3 3 3
Next is

1 ρ λ
| 70, 0+ >= χ φ + χλ φρ ψ00ρ
 2
+ χρ φρ + χλ φλ ψ00λ
 2 
(9.157)
2
To obtain u↑ , u↓ , d↑ , d↓ , we define the operator as in Equation 9.128:

u↑|70,0+ >p =<| 70, 0+ p ΣPu (3)P↑ (3) 70, 0+ p >

i

= 3 < 70, 0+ p |Pu (3)P↑ (3)| 70, 0+ p >


 

3
= [< χρ1 φλp + χλ1 φρp | Pu (3)P↑ (3) | χρ1 φλp + χλ1 φρp >
4 2 2 2 2

+ < χρ1 φρp − χλ1 φλp | Pu (3)P↑ (3) | χρ1 φρp − χλ1 φλp >]
2 2 2 2

All cross terms cancel, e.g.,


1 1
< φρp | Pu (3) | φλp >=< √ (udu − duu) | Pu (3) | − √ (udu + duu − 2uud) >= 0
2 6
Thus,
3
u↑|70,0+ >p = [< χρ1 | P↑ (3) | χρ1 >< φλp | Pu (3) | φλp >]
4 2 2

+ < χλ1 | P↑ (3) | χλ1 >< φρp | Pu (3) | φρp >


2 2

+ < χρ1 | P↑ (3) | χρ1 >< φρp | Pu (3) | φρp >


2 2

+ < χλ1 | P↑ (3) | χλ1 >< φλp | Pu (3) | φλp >


2 2

1 1
< φρp | Pu (3) | φρp >= √ (udu − duu) | Pu (3) | √ (udu − duu) >= 1
2 2
Hadron Physics 351
λ λ 1 1 1
< φp | Pu (3) | φp >=< √ (udu + duu − 2uud) | Pu (3) | √ (udu + duu − 2uud) >=
6 6 3

Similarly,
1
< φρ1 | P↑ (3) | χρ1 >= 1 and < χλ1 | P↑ (3) | χλ1 >=
2 2 2 2 3
 
3 1 1 1 1 4
u↑|70,0+ >p = 1× + ×1+1×1+ × = (9.158)
4 3 3 3 3 3
In the same manner,
3n o
u↓|70,0+ >p = < χρ1 | P↓ (3) | χρ1 >< φλp | Pu (3) | φλp >
4 2 2

+ < χλ1 | P↓ (3) | χλ1 >< φρp | Pu (3) | φρp >


2 2

+ < χρ1 | P↓ (3) | χρ1 >< φρp | Pu (3) | φρp >


2 2

+ < χλ1 | P↓ (3) | χλ1 >< φλp | Pu (3) | φλp >


2 2

1 1
< χρ1 | P↓ (3) | χρ1 >=< √ (↑↓↑ − ↓↑↑) | P↓ (3) |> √ (↑↓↑ − ↓↑↑) >= 0
2 2 2 2
1 2
< χλ1 | P↓ (3) | χλ1 >= (↑↓↑ + ↓↑↑ −2 ↑↑↓) | P↓ (3) | (↑↓↑ + ↓↑↑ −2 ↑↑↓) >=
2 2 6 3
Finally,
2
u↓|70,0+ >p = (9.159)
3
Next,
3
d↑|70,0+ >p = {< χρ1 | P↑ (3) | χρ1 >< φλp | Pd (3) | φλp >
4 2 2

+ < χλ1 | P↑ (3) | χλ1 >< φρp | Pd (3) | φρp >


2 2

+ < χρ1 | P↑ (3) | χρ1 >< φρp | Pd (3) | φρp >


2 2

+ < χλ1 | P↑ (3) | χλ1 >< φλp | Pd (3) | φλp >


2 2

2
< φρp | Pd (3) | φρp >= 0 and < φλp | Pd (3) | φλp >=
3
2
d↑|70,0+ >p = (9.160)
3
352 Group Theory in Particle, Nuclear, and Hadron Physics

One finds that,


1
d↓|70,0+ >p = (9.161)
3
We already calculated states with | 70, 2+ > when working out the de-
formed nucleon case. Next is the state | 20, 1+ >.
Now,
1 1 1
| 20, 1+ >= √ [φλ [χρ ψ1mA
2
]j= 2 − φρ [χλ ψ1mA
2
]j= 2 ] (9.162)
2

j= 21
q q
where [χX ψ1mA
2
]j 1 = 23 χX ψ 2 − 13 χX
− 12 11A
2
1 ψ10A
3= 2 2

j= 12
q q
and [χX ψ1mA
2
]j 1 = 13 χX

2
1 ψ10A −
2 X 2
3 χ 1 ψ1−1A
3 =− 2 2 2

( X=ρ or λ).

Therefore

u↑|20,1+ >p = 3 < p↑|20,1+ > | Pu (3)P↑ (3) | p↑|20,1+ > >

3 1 1
= [< φλp | Pu (3) | φλp >< [χρ ψ1mA
2
] 21 | P↑ (3) | [χρ ψ1mA
2
] 21 >
2 2 2

1 1
+ < φρp | Pu (3) | φρp >< [χλ ψ1mA
2
] 21 | P↑ (3) | [χλ ψ1mA
2
] 21 >]
2 2

1 1
< [χρ ψ1mA
2
] 21 | P↑ (3) | [χρ ψ1mA
2
] 21 >
2 2

3 1
= < χρ− 1 | P↑ (3) | χρ− 1 > + < χρ1 | P↑ (3) | χρ1
2 2 2 3 2 2

1 1
< χρ− 1 | P↑ (3) | χρ− 1 >=< √ (↑↓↓ − ↓↑↓) | P↑ (3) | √ (↑↓↓ − ↓↑↓) >= 0
2 2 2 2

1 1
< χρ1 | P↑ (3) | χρ1 >=< √ (↑↓↑ − ↓↑↑) | P↑ (3) | √ (↑↓↑ − ↓↑↑) >= 1
2 2 2 2

3 1 1
= ×0+ ×1=
2 3 3
Now,
1 1
< [χλ ψ|mA
2
] 21 | P↑ (3) | [χλ ψ|mA
2
] 21 >
2 2
Hadron Physics 353

2 1 2 2 1 1 5
= < χλ− 1 | P↑ (3) | χλ− 1 > + < χλ1 | P↑ (3) | χλ1 >= × + × =
3 2 2 3 2 2 3 3 3 3 9
As,
1
< χλ− 1 | P↑ (3) | χλ− 1 >= √ (↑↓↓ + ↓↑↓ −2 ↓↓↑) |
2 2 6
1 1 2
P↑ (3) | √ (↑↓↓ + ↓↑↓ −2 ↓↓↑) >= × 4 =
6 6 3
Thus,
3 1 1 5
u↑|20,1+ >p = [ × +1× ]=1 (9.163)
2 3 3 9
Next,
3 1 1
u↓|20,1+ >p = [< φλp | Pu (3) | φλp >< [χρ ψ|mA
2
] 21 | Pu (3) | [χρ ψ|mA
2
] 21 >
2 2 2

1 1
+ < φρp | Pu (3) | φρp >< [χλ ψ|mA
2
] 21 | Pu (3) | [χλ ψ|mA
2
] 21 >]
2 2

when
1 1 2
< [χρ ψ|mA
2
] 21 | P↓ (3) | [χρ ψ|mA
2
] 21 > < χρ− 1 | P↓ (3) | χρ− 1 >
2 2 3 2 2

1 2 1 2
+ < χρ1 | P↓ (3) | χρ1 >= × 1 + × 0 =
3 2 2 3 3 3
As,
1 1
< χρ− 1 | P↓ (3) | χρ− 1 >=< √ (↑↓↓ − ↓↑↓) | P↓ (3) | √ (↑↓↓ − ↓↑↓) >= 1
2 2 2 2
And,
1 1 2
< [χλ ψ|mA
2
] 21 | P↓ (3) | [χλ ψ1mA
2
] 21 > < χλ− 1 | P↓ (3) | χρ− 1 >
2 2 3 2 2

1 2 1 1 2 4
+ < χλ1 | P↓ (3) | χλ1 >= × + × =
3 2 2 3 3 3 3 9
As,
1 1
< χλ− 1 | P↓ (3) | χλ− 1 >= (↑↓↓ + ↓↑↓ −2 ↓↓↑) | (↑↓↓ + ↓↑↓ −0) =
2 2 6 3

3 1 2 4
u↓|20,1+ >p = [ × +1× ]=1 (9.164)
2 3 3 9
354 Group Theory in Particle, Nuclear, and Hadron Physics

TABLE 9.4: Results for u↑↓ and d↑↓ terms arising from
different parts of the configuration mixed quark model wave
functions
| 56, 0+ > | 560 , 0+ > | 70, 0+ > | 70, 2+ > | 20, 1+ >
↑ 5 5 4 2
u 3 3 3 3 1
u↓ 1
3
1
3
2
3
4
3 1
↑ 1 1 2 1 1
d 3 3 3 3 3
↓ 2 2 1 2 2
d 3 3 3 3 3

Next,
3  12  λ 2  12
d↑ = [< φλp | Pd (3) | χλp >< χρ ψ1mH
2

1 | p↑ (3) | χ ψ1mH 1 >
2 2 2

 12  λ 2  12 3 2 1
+[< φρp | Pd (3) | χρp >< χλ ψ1mH
2

1 | p↑ (3) | χ ψ1mH 1 >] = [ × + 0]
2 2 2 3 3

1
d↑|20,1+ >p = (9.165)
3
We also find,
2
d↓|20,1+ >p = (9.166)
3
We put all these result in Table 9.4.
Note as should be, for every case explicitly, u = u↑ + u↓ = 2 and d =
d + d↓ = 1.

Putting it all in the configuration mixed wave function we have


5 2 5 2 4 2 2 2
u↑ = a + b + c + d + e2
3 3 3 3
1 1 2 4
u↓ = a2 + b2 + c2 + d2 + e2
3 3 3 3
1 1 2 1 1
d↑ = a2 + b2 + c2 + d2 + e2
3 3 3 3 3
2 2 1 2 2
d↓ = a2 + b2 + c2 + d2 + e2 (9.167)
3 3 3 3 3
And
1
g1p =[15a2 + 15b2 + 9c2 − 9d2 − e2 ] (9.168)
54
1
g1n = [6c2 − 6d2 − 4e2 ] (9.169)
54
From the Bjorken sum rule,
Hadron Physics 355
TABLE 9.5: Polarized structure function values of different quantities
– the experimental values (E); the configuration mixed quark model values
(1 and 2, Equation 9.151); and the deformed nucleon results (D)
u↑ u↓ d↑ d↓ gA g1p g1n F
D Sz Lz
E - - - - 1.26 0,129 -0.031 0.632 0.176 0.324
1 1.66 0.34 0.35 0.65 1.59 0.271 +0.006 0.685 0.499 ∼ 0
2 1.65 0.35 0.34 0.66 1.61 0.273 +0.006 0.680 0.499 ∼ 0
D 1.46 0.53 0.33 0.66 1.26 0.187 -0.0225 0.582 0.297 0.203
.

gA
= 6(g1p − g1n ) (9.170)
gV
Using the Ellis-Jaffe sum rule,
F

gA 53 −1
g1p = (1 + F
D
) (9.171)
12 3 +1D
F

n gA 53 B −1
g1 = (−1 + F
 ) (9.172)
12 3 B +1
From these,
3
F = (4g p − g1n ) (9.173)
5 1
9
D= (2g p − 3g1n ) (9.174)
5 1
Sehgal [69] defines the spin of the proton as the sum of the total spin Sz
and the total orbital angular momentum Lz of its partons. With our picture,
1 9
Sz = (3F − D) = (g1p + g1n ) (9.175)
2 5
1
− Sz
L= (9.176)
2
In Table 9.5 we give all the relevant values of different physical quantities
evaluated here. We also show the corresponding experimental value [57] and
for the sake of comparison we also quote the deformed nucleon results [79],[80].

So these configuration mixed best wave functions (in Table 9.5, mix1 and
mix2 from [70] and [71], respectively) fail as to the g1p value. In addition, its
major failure is the g1n value, being very small in magnitude and positive in
sign, compared to its larger experimental value and empirically negative sign.
Also contrary to the experiment result, practically all of its spin resides in the
quarks degree of freedom only, exactly as in the non-relativistic quark model.
Deformed model [79],[80] succeeds where all others fail.
356 Group Theory in Particle, Nuclear, and Hadron Physics

9.7 Colour Confinement in QCD and Deformed Baryons


We saw that for the group SU (6)SF ⊗ SO(3)l ⊗ SU (3)C antisymmetry
for the three quarks in baryon, arises from SU (3)C space. In the rest of the
degrees of freedom, the state is fully symmetric, the so-called symmetric quark
model. So as per this picture, the QCD prediction is that the ground state
of the baryon is | 56, 0+ which is in the l = 0 orbital symmetric states. So
spherical symmetry of the low energy/ground states spin 21 and 32 baryon is a
clear prediction of QCD.
However we have seen above that the quark model, in order to fit a large
number of physical quantities, requires that the nucleon should be deformed in
the ground state itself. This is in direct conflict with the above QCD prediction
of sphericity. We treat this as an indication of the logical possibility that we
may be missing something in the quark model, or in QCD, or in both of these.
We note that the one-gluon exchange potential (OGEP) has a tensor term
which we suggested may induce the D-state admixture in the nucleon. We
shall see that the one-pion exchange potential (OPEP) too has a tensor term
which may contribute to the D-state in the nucleon. But it is well known that
this tensor term is too weak to induce the large PD ∼ 0.2 to 0.25 admixture
in the ground state. In fact the same OGEP also induces other configuration
mixtures (and not just the D-state) as we saw in the above configuration
mixed wave function. So what is so special about the D-state with respect to,
for example, the P -states. What distinguishes the D-state from the P -state
in the configuration mixture? Actually nothing, if OPEP and OGEP are the
only sources of inducing this configuration admixture. Some other source of
producing such large scale D-state admixture should be found. Now we discuss
one possible scenario which may explain this puzzle [81].
Note that in QCD,

3 ⊗ 3 ⊗ 3 = (3̄ ⊕ 6) ⊗ 3 → (3̄ ⊗ 3) ⊕ (6 ⊗ 3) → 1 ⊕ 8 ⊕ 8 ⊕ 10 (9.177)

So the colour singlet term necessarily arises from the 3 ⊗ 3̄ term. Each of
these colour states are sitting in the three-dimensional configuration space at
positions ‘i’ and ‘j’. Let us write this out in full,
X X X
φα (i)φα (j) = φα (i)φα (j) + φα (i)φα (j) (9.178)
i,j i6=j i=j

where ‘α’ stands for colour and i, j stand for the location of the colour fields
in space. The first term provides the well known colour state for meson con-
finement. We shall show here as to what the second term means.
Let us look at the colour singlet state of anticolour-colour,
Hadron Physics 357

3
1
2 3 1
2

(a) (b)

FIGURE 9.3: Locations of the three quarks at the positions 1, 2 and 3

1
φα (4)φα (3) ⇒ √ R̄(4)R(3) + B̄(4)B(3) + Ḡ(4)G(3)

(9.179)
3
Here anticolour is situated at position 4 and colour at position 3. The
reason for this somewhat odd numbering of positions will become clear below.
Using the identity in SU (3)c [28],[58],[82],

φα (4) = αβγ φβ (1)φγ (2) (9.180)


Clearly the position 4 should be the centre of mass of the positions 1 and
2 of identical quarks. Thus,

φα (4)φα (3) = αβγ φβ (1)φγ (2)φα (3) (9.181)


For baryons the positions 1, 2, 3 – at which the three colours are situated
and which are all independent – yield a totally antisymmetric state on the
exchange of any two colours at positions 1, 2, and 3. These three independent
positions 1, 2, and 3 are illustrated in Figure 9.3. Note that αβγ arises because
one has φα which leads to full antisymmetry in all the labels α, β, γ.
Note that the above antisymmetric state has a duality. Keeping the order
of the positions (123) fixed, there is an antisymmetry on the exchange of any
358 Group Theory in Particle, Nuclear, and Hadron Physics

pair of the three colours. Or we may keep the colours rigid and exchange any
two of the three positions, and antisymmetry still holds.
Next, we look at the colour singlet state given by the second term in
Equation 9.178. Let us write this as (labelling ‘3’ for the arbitrary location):
1 
√ R̄(3)R(3) + B̄(3)B(3) + Ḡ(3)G(3) (9.182)
3
Now using Equation 9.180 above, with antisymmetry in 1 ↔ 2 and the
third colour sitting fixed at position ‘3’,

1
√ [(B(1)G(2) − G(1)B(2)) R(3) + (G(1)R(2) − R(1)G(2)) B(3)
6
+ (R(1)B(2) − B(1)R(2)) G(3)] (9.183)
where the anticolour at position 3 creates a two-colour antisymmetric state
at positions 1 and 2, respectively, such that position 3 is the centre of mass
of these two. This is the point at which the third colour sits rigidly. This
situation is depicted in Figure 9.3(b). Position 3 is at the centre of mass while
position 1 and 2 can be exchanged. However in colour space, because of the
Levi-Civita tensor αβγ in Equation 9.181 there is still antisymmetry between
the exchange of any two colours. Thus the duality of the first antisymmetric
state is missing here. But the state is antisymmetric on the exchange of colours
nevertheless.
Note that this state is colour singlet and has a baryon number one as
well. As per the colour singlet hypothesis this colour singlet state for B = 1
should contribute to confinement. Here we claim that this is the colour singlet
confined state which has been missed so far for the confinement of baryons.
If r is the distance between particle 3 and particles 1 and 2, respectively
(Figure 9.3(b)), then clearly orbital terms like r2 , r4 , r6 ..... or even powers of
r will occur. These correspond to angular momentum states L = 2, 4, 6.....,
respectively. We note that for both J = 1/2 and J = 3/2 baryons these will
provide states of Regge trajectory and thereby form a natural explanation of
the same in this model.
Next note that L = 2 in proper combination with the spin degrees of the
three quarks will form the sought-after D-state admixture as indicated above.
Hence this provides a consistent mechanism for the D-state admixture of the
nucleons in the ground state and shows that, as per colour singlet confinement
hypothesis, there are actually two states, one provides the S-state plus the new
one giving the D-state admixture [81].
Hadron Physics 359

9.8 Solutions of Problems

Solution 9.1:
Normalization,
R ?
ψnlm ψnlm d~r = 1.

So,
R ∞ R 2π R π ?
r=0 φ=0 θ=0
(sinθdθ)dφ.ψ0d ψ0d r2 dr = 1

Using
R 2π Rπ ?
0
dφ 0 dθsinθYlm (θ, φ)Ylm (θ, φ) = δll0 δmm0

(Note in simpler cases one may need,


R R 2π Rπ Rπ

dΩ = φ=0
dφ θ=0
sinθdθ = 2π 0
sinθdθ = 4π)

Next,
q   12 2 R
3 ∞ 2 2
4 4α
α2 r2 e−α r r2 dr


15 π 0

Using [26],
R ∞ −px2 2n 1
1 Γ(n+ 2 )
0
e x dx = 2 p(n+ 12 )

get, √
R∞ 2 2
15 π
0
r6 e−α r
dr = 16 α7

This finally gives 1 for the above integral.

————————————————–
Solution 9.2:
√ →

d
Note p~1 − p~2 = m dt (r~1 − r~2 ) = 2m ddtρ

Thus p21 + p22 − 2p~1 · p~2 = 2m2 ( d~


ρ 2
dt ) = 2pρ
2

d

and p~1 + p~2 − 2p~3 = m dt (r~1 + r~2 − 2r~3 ) = 6~

Thus p~1 2 + p~2 2 − 2~


p.p~2 + 4p23 − 4p3 .(p~1 + p~2 ) = 6pλ 2

d
next p~1 + p~2 + p~3 = m dt d
(r~1 + r~2 − r~3 ) = 3m dt ~ = PCM
RCM ~
360 Group Theory in Particle, Nuclear, and Hadron Physics
p~2ρ p~ ~
2
PCM p~1 2 2
So 2m + λ2
2m + + p2m
2M
~2
=+ 2mp~3
2m
 2  → 2 


r1√−→

r2 −
r1 +→−
r√2 −2→−
r3
and 23 Kρ2 + 32 Kλ2 = 32 K 2
+ 6
→−2 →−2 → −2
= K r1 + r2 + r3 − 2→ −r1 · →

r2 − 2→ −
r1 .→

r3 − 2→−
r2 .→−

r3

= 21 K |→

r1 − →

r2 | + 12 K |→

r1 − →−
r3 | + 12 K |→

r2 − →

2 2 2
r3 |
n 2
p2λ
o
~ 2 pρ
Thus H = pCM 3
2M + 2m + 2m + 2 kρ + 2 kλ
2 3 2

2
H = pCM
~
2M + H0
————————————————–
Solution 9.3:
q
8 α5
2
 2 2 2 2
 −α2 (ρ2 +λ2 )/2
ψ2mS = 15 π ρ Ym (Ωρ ) + λ Ym (Ωλ ) e

8 α10
ψ ? ψd3 ρd3 λ =
R
15 π 2
RR h 2 2 2 2 2
i
(Ωλ )e−α (ρ +λ ) d3 ρd3 λ
2 2
ρ4 Ym (Ωρ ) + λ4 Ym (Ωλ ) + 2ρ2 λ2 Ym
2 2
(Ωρ )Ym

The last term over spherical harmonics is zero and the first two are similar,

8 α10
2 2 2 2
ρ4 Ym2 (Ωρ ) e−α (ρ +λ ) d3 ρd3 λ
RR
= 15 π 2 .2

8 α10
R∞ 2 2 R∞ 2
λ2
R 2
= 15 π 2 · 2 · 4π. 0
ρ6 e−α ρ
dρ 0
λ2 e−α dλ Ym2 (Ωρ ) dΩρ
√ √
8 α10 π 1 π
= 15 π 2 · 2 · 4π. 15
16 α7 . 4 α3 =1

————————————————–
Solution 9.4:
m1 = m2 = m, m3 = ms . As in ρ − case above,
1  <US (r)>M <US (r)>M
m(K ? ) − m(K) = 4 − (− 34 ) mms = mms

Hence,

m
(m(K ? ) − m(K)) / (m(ρ) − m(π)) = ms

For Σ, Σ? ,


S~1 ·S~2 S~0 .S3
< USS (r) >B =< m2 + mms >< Us (r) >B

When S~0 = S~1 + S~2 , S


~ = S~0 + S~3 So
Hadron Physics 361

< USS (r) >B = 12 S 0 (S 0 + 1) − 2 · 34 · <USm(r)>


 B
2

n o
+ 12 S (S + 1) − S 0 (S 0 + 1) − 34 . <Umm
S (r)>B
s

S 0 = 1 as in Σ? ,Σ, u- and d-quarks are in flavour symmetric state and so


to obtain the fully symmetric flavor-spin state S 0 = 1.

(m(Σ? ) − m(Σ) = 12 ( 23 · 5
2 − 34 ) <Umm
S (r)>B
s

m
So, m(Σ? ) − m(Σ)/(m(∆) − m(N )) = m s
∼ 300/450 ∼ 23 which is a good
? ?
fit. Next (m(Σ ) − m(Σ)) / (m(K ) − m(K)) = (m(∆) − m(N )) / (m(ρ) − m(π)) ∼
0.5, which agrees with the experimental value.
————————————————–
Solution 9.5:
The time-independent Schroedinger equation in three dimensions,

~2 ~ 2
− 2m ∇ u + V u = Eu

Taking u(r, θ, φ) = R(r)Y (θ, φ) separates out the angular and the radial
equations.

h2 1 d 2 dR λ
 
− 2m r 2 dr r dr + V + r2 R = ER

h2 1 ∂ ∂V ~2 1 ∂2Y

− 2m sin θ ∂θ sin θ ∂θ − 2m sin2 θ ∂φ2 = λY

The Y (θ, d) solution represents the angular part of the wave function for
any spherically symmetric (central) potential V(r).

One solves the angular equation and obtain for

λ0 = 2mλ
~2 = l (l + 1)

We get

h2 1 d 2 dR
 h l(l+1)~2
i
− 2m 2
r dr r dr + V (r) + 2mr 2 R = ER

Substitute χ(r) = rR(r) to get,


h i
2
h2 d χ l(l+1)~2
− 2m dr 2 + V (r) + 2mr 2 χ = Eχ

For the deuteron we have a central potential which is attractive V ≤ 0.


The effective potential is given above in the square bracket . The second term
in the square bracket is a repulsive term and thus l = 0 provides the most
attraction. Hence in the ground, to obtain the maximum attraction, the orbital
362 Group Theory in Particle, Nuclear, and Hadron Physics

angular momentum must be zero. So for the central interaction, ground state
of the system should be spherical.
————————————————–
Solution 9.6:
The quadrupole moment operator Q ∼ r2 Y2 (θ, d) varies as the L = 2 or-
bital angular momentum. The quadrupole moment is given as the expectation
value < j |Q| j >. Now for deuteron having spin j = 1, ~1 + ~2 ∼ ~1, and a non-
zero value of the quadrupole moment is possible (and which is also observed).
But for spin j= 21 , ~12 + ~2 6= ~21 and thus the above expectation value is zero.
Hence a nucleon and any nucleus with spin 21 , does not have non-zero value
of the quadrupole moment.
————————————————–
Solution 9.7:
(a)
<n↑ | σ+ (i)ν− (i)|Σ− ↓ >
P
gA = <n↑ |Σν− (i)|Σ−↑ >
i

n↑S | σ+ (i)ν− (i)| Σ− ↓


P
<
n S > o
ρ ↑ ↓
= 2 < φn |v− (3)| φ − >< χρ |σ+ (3)| χ↓
3 ρ ↑ λ λ
ρ > + < φn |v− (3)| φΣ− >< χλ |σ+ (3)| vλ >
Σ

< φρn |v− (3)| φρΣ− >= √12 (udd − dud) |v− (3)| − √12 (dsd − sdd) >= 0

< φλn |v− (3)| φλΣ− >= 0

So < n↑S | σ+ (i)v− (i)| Σ− ↓ 3 2


− 13 = − 13
P  
S > 2 0×1+ 3

< n↑S | v− (i)| Σ−↑


P
S >= 1

< n↑D | σ+ (i)v− (i)| Σ− ↓ 3


0 × (− 13 ) + 2
− 13 = − 13
P  
D > 2 3

< n↑D |Σν− (i)| Σ−↑


D >= 1

(1−PD )(− 31 )+PD (− 13 )


gA (Σ− → ne− ν̄e ) = (1−PD )+PD .1 = − 13

Note no modification due to the D-state.

(b)
<Σ+↑ | i σ+ (i)v− (i)|Ξ0↓ >
P
gA = <Σ+↑ |Σv− (i)|Ξ0↑ >

(1−PD ). 53 +PD (− 31 )
= (1−PD ).1+PD .1

gA Ξ0 → Σ+ e− ν̄e = 5

3 − 2PD
Hadron Physics 363

<Σ0↑ σ+ (i)v− (i) Ξ− ↓ >
P
i P
(c) gA = <Σ0↑ | v− (i)|Ξ−↑ >
 
(1−PD )· 35 . √12 +PD − 13 · √12
= (1−PD ) √12 +PD . √12

gA Ξ− → Σ0 e− ν̄e = 5

3 − 2PD
−↓
<Ξ0↑ | σ
P
+ (i)τ+ (i)|Ξ >
(d) gA = <Ξ0↑ | τ+ (i)|Ξ−↑ >
P

(1−PD )(− 13 )+PD (− 13 )


= (1−PD ).1+PD .1

gA Ξ− → Ξ0 e− ν̄e = − 13


   
1 1
(1−PD ) − 3√ +PD − 3√

= − 13
0
 2 2
(e) gA Σ → pe ν̄e = (1−PD ) √12 +PD √12
√  √ 
(1−PD ). 2 3 2 +PD − 32
(f) gA Σ0 → Σ+ e− ν̄e = 2


(1−PD ) 2+PD 2
√ = 3 − PD

(g) gA (Σ− → Λe− ν̄e )


<Λ↑ | σ −↓
P
+ (i)τ+ (i)|Σ >
= <Λ↑ | τ+ (i)|Σ−↑ >
P

Find < φρΛ |τ+ (3)| φρΣ− >= √1


6

and < φλΛ |τ+ (3)| φλΣ− >= − √16

and so in the denominator both,

< Λ↑S | τ+ (i)| Σ− ↑ ↑ −↑


P
S >=< ΛD |τ+ (i)| ΣD >= 0

So gA → ∞ and hence get G


GA = 0
V

————————————————–
Solution 9.8:

3 ρ Q(3) ρ
< Σ+ +
φ + >< χρ↑ |σz (3)| χρ↑ >

S |µz | ΣS >= [< φ Σ+
2 mq (3) Σ

λ Q(3) λ e~
φ + >< χλ↑ |σz (3)| χλ↑ >]

+ < φΣ+
mq (3) Σ 2c
 
8 1 e~
= +
9mu 9ms 2c
As D− part is zero (taking the general result from above),
 
8 1 e~
< Σ+ |µ| Σ+ >= (1 − PD ) −
9mu 9ms 2c
364 Group Theory in Particle, Nuclear, and Hadron Physics

————————————————–
Solution 9.9:
For D states,

1  21  S 2  12
< n↑D | Nd↑ | n↑D >= [< φρn χS ψ2mρ
2 ρ

1 |Nu↑ | φn χ ψ2mρ 1 >
2 2 2

 12  S 2  12
+ < φλn χS ψ2mλ
2 λ

1 |Nu↑ | φn χ ψ2mλ 1 >]
2 2

Now,
 12  S 2  12
< φρn χS ψ2mλ
2 ρ

1 |Nu↑ | φn χ ψ2mλ 1 >
2 2

1 1
= < φρn χS3 |N u↑ | φρn χS3 > + < φρn χS1 |N u↑ | φρn χS1 >
10 2 2 5 2 2

3 2
+ < φρn χS− 1 |N u↑ | φρn χS− 1 > + < φρn χS− 3 |N u↑ | φρn χS− 3 >
10 2 2 5 2 2

1 1
< φρn χS3 |N u↑ | φρn χS3 >=< √ (udd − dud) ↑↑↑ k √ (udd − dud) ↑↑↑>= 1
2 2 2 2

1 1
< φρn χS1 |N u↑ | φρn χS1 >=< √ (udd − dud) √ (↑↑↓ + ↑↓↑ + ↓↑↑)
2 2 2 3

1 2
| √ udd (↑↑↓ + ↑↓↑ +0) − dud (↑↑↓ +0+ ↓↑↑) >=
2 3

1 1
< φρn χS− 1 |N u↑ | φρn χS− 1 >=< √ (udd − dud) √ (↓↓↑ + ↓↑↓ + ↑↓↓)
2 2 2 3

1 1
| √ udd (0 + 0+ ↑↓↓) − dud (0+ ↓↑↓ +0) >=
6 3
Therefore,
 21  S 2  12 1
< φρn χS ψ2mλ
2 ρ

1 |Nu↑ | φn χ ψ2mλ 1 >=
2 2 3
So,

< n↑ | Nu↑ | n↑ >=< (1 − PD ) < n↑S | Nu↑ | n↑S > +PD < n↑S | Nu↑ | n↑S >

1 1
= (1 − PD ) × + PD ×
3 3
Hadron Physics 365

Finally,
1
u↑n =
3

1  12  S 2  12
< n↑D | Nd↑ | n↑D >= [< φρn χS ψ2mρ
2 ρ

1 |Nd↑ | φn χ ψ2mρ 1 >
2 2 2

 12  S 2  12
+ < φλn χS ψ2mλ
2 λ

1 |Nd↑ | φn χ ψ2mλ 1 >]
2 2

 
1 2 2 2
= + =
2 3 3 3
Therefore,
5 2 5
+ PD × = − PD d↑n = (1 − PD )
3 3 3
————————————————–
Solution 9.10:

Z 2π Z π
dφ sin θdθ = 4π
φ=0 0

∞ √
1 τ 32
Z 
2 2 1 π
ρ2 e−α ρ
dρ = =
0 2 (α2 ) 32 4 α3

∞ λ
α3
ZZ Z Z
? 2 2 2
2 λ2
0
ψ00S 0
ψ00S d3 ρd3 λ = ( 3 )2 (4π) [ ρ2 e−α ρ
dρ λ2 e−α dλ]
π 2 0 0

 √  √ 
α6 1 π 1 π
= .16π 2 =1
π3 4 α3 4 α3
ZZ
?
2 2
ψ00S ψ00S d3 ρd3 λ

3 Z ∞ ∞
α3
 Z 
1 2  2 2 2
 −α2 (ρ2 +λ2 )/2 2 2
= (4π) 3 3−α ρ +λ e ρ λ dρdλ
3 π2 0 0

Z ∞ Z ∞ Z ∞ Z ∞
2 −α2 ρ2 2 −α2 λ2 4 −α2 ρ2 2
λ2
[...] = 9 ρ e dρ λ e dλ+α 4
ρ e dρ λ2 e−α dλ
0 0 0 0

Z ∞ Z ∞ Z ∞ Z ∞
2 2 2
λ2 2 2 2
λ2
+α4 ρ2 e−α ρ
dρ λ4 e−α dλ + 2α4 ρ4 e−α ρ
dρ λ4 e−α dλ
0 0 0 0
366 Group Theory in Particle, Nuclear, and Hadron Physics

Z ∞ Z ∞ Z ∞ Z ∞
2 2 2
λ2 2 2 2
λ2
−6α2 ρ4 e−α ρ
dρ λ2 e−α dλ − 6α2 ρ2 e−α ρ
dρ λ4 e−α dλ
0 0 0 0

So
16π 2 α6 9
 
15 15 9 9 9 π
= + + + − − =1 (9.184)
3 π 3 10 64 64 32 16 10 α6
————————————————–
Chapter 10
Glashow-Salam-Weinberg Model

10.1 Its Chiral and Non-Chiral Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367


10.2 Spontaneous Symmetry Breaking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 372
10.3 GSW Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 374
10.4 Millicharged Neutrinos . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 380
10.5 Second Generation Fermions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 381
10.6 Right-Handed Neutrino . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 383
10.7 Gauge Bosons and Neutral Currents . . . . . . . . . . . . . . . . . . . . . . . . . . . . 386
10.8 Neutrino Oscillation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 389
10.9 Appendix F: Lorentz and Poincare Groups . . . . . . . . . . . . . . . . . . . . . . 399
10.10 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 407

10.1 Its Chiral and Non-Chiral Structure


In the late 1950s and up to ∼1961 the theory of the strong interaction based
on the SU (2)I isospin symmetry group was extended to the group SU (2)I
⊗U (1)Y . This was performed in order to incorporate the newly discovered
concept of strangeness. As we saw in Chapter 6, in order to fit all the charges of
the known hadrons, the Gell-Mann-Nishjima relation was successfully utilized
for the group SU (2)I ⊗U (1)Y . This is defined as:
Y
Q = I3 + (10.1)
2
where I3 is the diagonal generator of the group SU (2)I and Y is the hyper-
charge, Y=B+S, a quantum number associated with the group U (1)Y ; while
B and S are baryon number and strangeness quantum numbers, respectively.
The larger group SU (2)F with (u, d, s) quarks was yet to be proposed (in
∼1963).
So it was but natural that, at that time, a unified theory of the weak and
the electromagnetic interaction was built around a similar group structure
SU (2)L ⊗ U (1)YW . This is what was actually suggested by Glashow in 1961.
Here SU (2)L is a left-handed group utilized to account for parity violation in
the weak interaction and U (1)YW is a weak-hypercharge YW group. Both of
these are chiral groups, in contrast to the strong interaction group SU (2)I ⊗
U (1)Y , which is completely non-chiral in nature. Thus it was logical to define

367
368 Group Theory in Particle, Nuclear, and Hadron Physics

the electric charge for the electromagnetic-weak group SU (2)L ⊗ U (1)YW as


having analogous structure as Equation 10.1 above [[46],p.309],
YW
Q = (I3 )L + (10.2)
2
Below we investigate two fundamental aspects of the structure of the elec-
tric charge in the above electro-weak model.
(A) The most astonishing aspect of Equation 10.1 for the electric charge
Q is that the quantity on the LHS is non-chiral in nature, while both the
quantities on the RHS are entirely chiral. However no such problem exists
for the corresponding equation of electric charge in the strong interaction,
where both the LHS and the RHS are non-chiral. Let us see in what manner
Equation 10.2 may be considered as yielding a non-chiral electric charge, while
composed of a linear combination of chiral charges.
First we shall attempt to comprehend why the SU (2)L group in itself is
not sufficient to unify the electromagnetic and weak interactions, and why a
minimal SU (2)L ⊗U (1)YW is required to accomplish this task. Note that the
two-flavour (ν̄e , e− )L or (u, d)L left-handed weakly charged currents and its
conjugate may be associated with two of the three generators (T + , T − ) of the
SU (2)L group. Can we thus identify the neutral non-chiral electromagnetic
current with the third generator T3 of the SU (2)L group? Note that for the
(ν¯e , e− ) pair the left-handed charge-raising current is
1
Jµ+ =
[ν¯e γµ (1 − γ5 )e]
2
and the charge lowering current is Jµ− . Let us write the two charges Q+ (t)
and Q− (t) of the SU (2)L group as

Z Z
1 †
Q+ (t) = d3 yJ0+ (r) = d3 xνe† (1 − γ5 )e; Q− (t) = Q+ (10.3)
2
Using the equal-time commutation relation for fermions,

[ψk† (~x, t), ψl (~x0 , t)]+ = δkl δ 3 (~x − ~x0 ) (10.4)


and thus,

Z
− 1
+ 3
[Q (t), Q (t)] = 2Q (t); Q (t) = 3
d3 x[νe† (1 − γ5 )νe − e† (1 − γ5 )e] (10.5)
4

Note that here Q3 (t) is the left-handed neutral chiral current. It is not the
non-chiral Qem which is say Qem = d3 xe† e. So Q+ , Q− , Q3 form the closed
R

SU (2)L algebra from which Qem has been left out. Thus we clearly have
to enlarge the SU (2)L group to incorporate the non-chiral electromagnetic
current. Hence one more generator, using the principle of minimality, should
Glashow-Salam-Weinberg Model 369

be added to the three already present. Thus minimum enlargement suggests


the SU (2)L ⊗ U (1)YW group which is chiral in structure. How do we obtain
the non-chiral charge from here?
Now the problem is that, for the GSW model with the group SU (2)L ⊗
U (1)YW , the charge is entirely chiral in nature. How can we derive a non-chiral
charge of QED from the above apparently incongruent combination?
To obtain a non-chiral U (1)em group we require a non-linear mechanism
of spontaneous symmetry breaking by an Englert-Brout field (weak isospin
doublet). Thus U (1)em , which is an unbroken non-chiral field, is generated
to provide a linear combination of the chiral fields of the SU (2)L ⊗ U (1)YW
group as defined in Equation 10.2. Note the significance of the spontaneous
symmetry breaking mechanism to create a non-chiral unbroken structure of
the broken SU (2)L ⊗ U (1)YW field [[31] p.22, p.335].
To explain the maximal parity violation in the weak interaction, the uni-
versal (V-A) theory was put forward by Sudarshan and Marshak in 1957 [83].
This theory exploited the principle of “chirality invariance”. The principle of
chirality invariance takes its cue from the two-component theory of neutrinos,
which states that the massless neutrino satisfies the γ5 -invariant Weyl equa-
tion and exhibits maximal parity violation. This “neutrino paradigm”, as it
was termed, was generalized to massive fermions partaking in the weak inter-
action. The new chirality invariance principle, now demanding that all lepton
and baryon weak charged currents, should be invariant under the individual
replacement of each spinor wave function φi by its chirality transform γ5 φk
or φk → γ5 φk [31].
Let θ represent all possible relativistically covariant combinations of
the Dirac γµ operators. There are only five: (taking, γ5 = −iγ0 γ1 γ2 γ3 ):
1(S), γ5 (P S), γµ (V ), γ5 γµ (A), 21 [γµ , γν ](T ).
Then “chirality invariance” demands that φn θφl (k 6= l) for charged two-
flavour current. Then its general Lorentz structure should be θ = γµ (1 − γ5 ),
i.e., (V-A) form. Thus the Lorentz structure of the charged fermion (lepton
and baryon) current ensures that only left-handed chiral fermions be involved
in charged currents.
Next let us invoke the same principle of chirality invariance for neutral
weak (chiral) and electromagnetic currents (non-chiral). It is necessary to do
so as the neutral weak current is anyway a generator of the SU (2)L group,
as demonstrated above. Note that in contrast to the weak charged currents
which are two-flavour in nature, neutral weak and electromagnetic currents are
one-flavour in nature. This difference leads to the condition that the neutral
fermion currents possess the Lorentz structure of (V+a A) form, where ‘a’
is a constant. The principle chiral invariance as applied to neutral currents
ensures a consistency with the parity-conserving (V) neutral electromagnetic
current along with the form of the electro-weak neutral current.
Now we can see how in Equation 10.2, the linear combination of chiral
charges on the right may be equated with the non-chiral charge on the left.
It is the existence of a common principle of chirality invariance [[31], p. 22,
370 Group Theory in Particle, Nuclear, and Hadron Physics

p. 335] in the parity-violating (chiral) SU (2)L ⊗ U (1)YW weak group and


also in the parity-conserving (non-chiral) electromagnetic group U (1)em . This
allows for the chiral combination of charges in SU (2)L ⊗ U (1)YW and it is the
spontaneous breaking of this chiral SU (2)L ⊗U (1)YW group into the non-chiral
U (1)em group which establishes this connection (as in Equation 10.2). Thus
the Englert-Brout mechanism clearly plays a fundamental role in creating this
unique relationship between chiral and non-chiral charges in the electro-weak
symmetry.
The above argument by Marshak is merely a consistency argument. It just
establishes that the SSB by an Englert-Brout field permits that the two chi-
ral charges add up to provide a non-chiral charge. The photon field created
thereby is not only massless, it is also non-chiral in nature. What this im-
plies is that it couples identically to the left-hand and the right-hand charges:
meaning QL = QR . Now this last condition is not necessarily an unequivocal
requirement of the SSB by a scalar field, but is solely an additional constraint.
Only the existence of charges satisfying the above condition will ensure the
non-chiral nature of the full U (1)em , i.e., Quantum Electrodynamics. We shall
see below that this is indeed the case.
(B) In the Glashow-Salam-Weinberg (GSW) model of particle physics,
which is based on the group structure SU (2)L ⊗ U (1)YW , all the known ele-
mentary particles are classified in terms of three independent “generations”.
These generations actually form a repetitive structure. In the GSW model
the first generation consists of quark-lepton doublets {(νe , e− )L , (u, d)L },
while the second and the third generations consist of {(νµ , µ− )L , (c, s)L } and
{(ντ , τ − )L , (t, b)L } pairs, respectively. Thus one may confine oneself to study-
ing any one, say the first generation only.
Now the charges of all the members of the first generation are obtained by
picking up values of YW so as to obtain the correct charges for the doublet,
e.g., YW = −1, which produces the correct charges of (νe , e− )L using Equation
10.2 and YW = + 31 to yield the correct charges for the (u, d)L pair.
Now for the strong interaction within the group SU (2)I ⊗U (1)Y , the defini-
tion of the electric charge as provided in Equation 10.1 is not understandable.
Because we do not comprehend as to what maintains the correlation between
I3 and Y in Equation 10.1. This is because the isospin and the hypercharges
are independent of each other. However when the strong group was extended
to SU (3)F ⊃ SU (2)L ⊗ U (1)Y in 1963, then the definition of the electric
charge became fully consistent. This occurred as the two generators in the
electric charge of SU (3)F are identified by the two diagonal generators I3 and
Y. And thus this definition of electric charge is consistently achieved entirely
within the structure of the SU (3)F group itself.
However, such an embedding of the electro-weak group SU (2)L ⊗ U (1)YW
into a larger group does not happen and the arbitrariness in the definition
of the corresponding electric charges in Equation 10.2 remains. Hence the
construction of the electric charge, as provided above in Equation 10.2, is
quite discretionary and subjective in nature. Because of this, in fact, another
Glashow-Salam-Weinberg Model 371

equally popular definition of the electric charge in the SU (2)L ⊗U (1)YW group
exists in the literature. This is as follows [6]:

Q = (T3 )L + Yw0 (10.6)


We distinguish Yw0
as a generator of the group U (1)YW from Yw used in
Equation 10.2. We may do so and still obtain the correct electric charges for
0
the first generation pair with YW = − 21 and Yw0 = 16 for the lepton and the
quark pairs, respectively.
Hence, it has been commonly felt – both on account of the arbitrary con-
struction of the electric charge, as well as the hypercharge quantum numbers
being whimsically put in by hand – that no consistent charge quantisation
exists in the GSW model.
This lack of a consistent understanding of the quantisation of the electric
charge is popularly considered to be fundamental weakness of the GSW model.
However, notice that the hypercharges in Equations 10.2 and 10.6 are
related by a common factor of 2. Thus we expect that there should exist a
hidden parameter present in the electric charge.
The following definition in terms of an unknown parameter Yφ suggests
itself:
1
Q = (T3 )L + ( )YW (10.7)
2Yφ
Yφ = 1 and Yφ = 12 yields the two above definitions of the electric charge
in Equations 10.2 and 10.6, respectively.
If we accept this new definition of the electric charge in Equations 10.7
and assume that the matter field (quarks and leptons) hypercharges are pro-
portional to Yφ , then all the electric charges are properly defined for the first
generation as below for the group SU (2)L ⊗ U (1)YW .

 
u Yφ 1 1 Yφ 2 1
; Y = → Q(u)L = + · = ; Q(d)L = −
d L 3 2 2Y φ 3 3 3

4 1 4 2
uR ; Y = Yφ → Q(u)R = 0 + · Yφ =
3 2Yφ 3 3
2 1 2 1
dR ; Y = − Yφ → Q(d)R = 0 + (− Yφ ) = −
3 2Yφ 3 3

 
ve 1 1
; Y = −Yφ → Q(ν)L = + (−Yφ ) = 0, Q(e)L = −1
e− L 2 2Yφ

1
eR ; Y = −2Yφ , Q(eR ) = 0 + (−2Yφ ) = −1 (10.8)
2Yφ
Thus the electric charge is consistently quantized in the GSW model.
372 Group Theory in Particle, Nuclear, and Hadron Physics

But what is this new unknown Yφ that we have introduced here? Now
around 1961, when these charges of Equations 10.2 and 10.6 were introduced,
there was no other field present in the theory other than the matter fields as
provided in Equation 10.8. But in 1964 Englert-Brout suggested a new field
that would break the corresponding symmetry spontaneously, and thereafter
Higgs proposed an associated boson with this new field. This Englert-Brout
field was later used by Weinberg and Salam in 1967 to generate gauge boson
masses in their model. Thus there is an extra Englert-Brout field in addition
to the above field in Equation 10.8. What are the charges of this new field?
They, using Equation 10.1 and the Englert-Brout field hypercharges Yφ = 1,
obtained the charges of this field. Yφ will be 21 if they were to use Equation
10.6 for the charges. But now we know that these arbitrary values of Yφ are
not required. Using Equation 10.7 and utilizing Yw = Yφ by itself is enough
to obtain the correct charges for the Englert-Brout field. We will see more of
this below.
Hence the proper definition of electric charge in the GSW model is given in
Equation 10.7, and this correctly specifies the quantized electric charges of the
Englert-Brout field and all the known matter-particle charges. This Equation
10.7 is clearly consistent with the earlier arbitrary constructions of the electric
charge for special values of Yφ (Equations 10.2 and 10.6).
The above approach in deriving fully quantized electric charges in the
GSW model may be labelled as phenomenological. However what has been
accomplished here on the basis of consistency arguments within the full GSW
model, including the Englert-Brout field, will be demonstrated below to be
derived on the basis of its complete symmetry structure inherent in the GSW
model group of SU (2)L ⊗ U (1)YW .

10.2 Spontaneous Symmetry Breaking


The group structure of the GSW model was already outlined by Glashow
in 1961. However the gauge particles mediating the weak interaction were left
massless. Salam and Weinberg independently, in 1967, generated masses for
these gauge bosons through a mechanism of spontaneous breaking of the gauge
symmetries. Veltman and ’t Hooft demonstrated in 1971/72 that, in spite of
spontaneous breaking of the symmetry, the theory remains renormalizable.
The mechanism of spontaneous symmetry breaking was first suggested by
F. Englert and R. Brout in August 1964 [84]. This was subsequently followed
by additional papers from P. Higgs in October 1964 [85] and G. Grualnik,
C. R. Hagen and T.W.B. Kibble in November 1964 [86].
To clarify the situation and to give proper credit, we should clearly under-
stand that the issue of spontaneous symmetry breaking in the GSW model
involves two distinct aspects – one is the mathematical mechanism of sponta-
Glashow-Salam-Weinberg Model 373

neously breaking the symmetry, and the other is the existence of an associated
massive scalar boson.
Let us understand as to who had incorporated what degrees of freedom
in their models. As pointed out by Ellis, Gaillard and Nanopoulos [87], the
Englert-Brout mechanism possessed no scalar bosons in its framework of spon-
taneous symmetry breaking of non-Abelian gauge theory. Later Higgs was the
first one to demand a massive scalar boson in his Abelian group study. Still
later, Guralnik-Hagan-Kibble introduced a massless scalar in their Abelian
group discussion.
So we have to comprehend as to what is the role of a massive scalar boson
in the actual mechanism of spontaneous symmetry breaking within the GSW
model.
For clarity’s sake, let us quote directly from a historical survey of the
spontaneous symmetry breaking by Brout and Englert (1998) [88]:
1. “Massive gauge vector bosons are an inevitable consequence of SBS,
independently of the dynamical mechanism which causes the breaking, scalar
fields, bound state condensate. . . ”
2. “Massive scalars occur in channels orthogonal to the NG channels. They
occur in any global SBS and are thus not a specific feature of mass generations
for gauge field: their physics is accordingly much more sensitive to dynamical
assumption.”
3. “In all cases SBS in gauge theories is characterized by NG bosons which
are ‘eaten up’ by the YM fields, giving them longitudinal polarisation and
mass. It is this phenomenon which allows for consistent renormalizable theories
of massive gauge vector mesons. . . ”
(acronyms above: SBS: spontaneously broken symmetry, NG: Nambu-
Goldstone, YM: Yang-Mills)
Thus the SBS clearly provides a renormalizable structure to the GSW
model which is independent of the existence of a scalar meson or boson. Hence
the scalar boson or meson is merely an addendum to the SBS mechanism.
Consequently we shall term the mechanism of spontaneous symmetry breaking
as the Englert-Brout mechanism (EB mechanism), as they were definitely the
first ones to have proposed this mechanism for a non-Abelian gauge group,
without invoking a scalar boson or meson. We shall denote the subsequent
addendum of a scalar boson or meson as the Higgs boson, giving credit to the
fact that Higgs was the first one to demand a massive boson in addition to
SBS of his Abelian U(1) group.
Our convention of distinguishing the Englert-Brout mechanism from the
Higgs boson is consistent with the citation of the Nobel Prize in Physics in
2013 to F. Englert and P. Higgs (R. Brout, unfortunately having passed away),
which states, “For the theoretical discovery of a mechanism that contributes
to our understanding of the origin of mass of subatomic particles, and which
recently was confirmed through the discovery of the predicted fundamental
particle by the ATLAS and CMS experiments at CERN’s large Hadron Col-
lider.”
374 Group Theory in Particle, Nuclear, and Hadron Physics

However, the alert reader may take issue with the word “predicted”, as we
now clearly understand that the Higgs boson is not an unequivocal prediction
of the spontaneously broken symmetry of the Englert-Brout mechanism. The
Higgs boson is actually an independent addendum to the EB mechanism. The
EB mechanism may exist independently of the Higgs boson, but the latter
depends upon the validity of the EB mechanism on an a priori basis.

10.3 GSW Model


In the Glashow-Salam-Weinberg (GSW) model the symmetry (to be de-
fined below) spontaneously broken by an Englert-Brout field is taken to be an
SU (2)L group doublet:
 +
φ
φ= (10.9)
φ0
where the superscripts yield the charges of these complex fields,
φ1 + iφ2
φ+ = √
2
φ3 + iφ4
φ0 = √ (10.10)
2
Salam and Weinberg using the Glashow expression for the electric charge,
i.e., Equation 10.2, took the hypercharge for the Englert-Brout field for
SU (2)L ⊗ U (1)YW as Yφ = 1. This gave them the electric charges for the
Englert-Brout doublet field. But this was an arbitrary step, as suppose if they
were to use the other popular definition – i.e., Equation 10.6 – for the charge,
then they would have had to take Y φ = 12 to obtain the proper charges.
The above freedom to select the electric charges in the Glashow-Salam-
Weinberg model was taken to indicate that the electric charge was not properly
quantized in it. The issue of charge quantization means for example, why the
d-quark charge is 13 of the electronic charge in this framework. If this was
all, then obviously the electric charge would not be considered to be properly
quantized in the GSW model.
This absence of electric charge quantization was considered to be a ma-
jor shortcoming of the GSW model. In fact this perceived deficiency was the
primary motivation for studying models which extended beyond the GSW
framework. These were given the fanciful name of ‘Grand Unified Theories’
(GUTs). In 1974 a GUT called SU (5) was constructed by Georgi and Glashow.
The proclaimed primary success of that SU (5) model was that charge quanti-
zation was built into it. It was asserted that this implied that GUTs were on
the right track, as they exhibit electric charge quantization, which the GSW
Glashow-Salam-Weinberg Model 375

model does not. However, the SU (5) GUT model simultaneously predicted
that the proton would decay with a lifetime of ∼ 1031 years.
Great excitement was hence occasioned when the above proton decay pre-
diction of SU (5) was supposedly confirmed in 1982 by an Indo-Japanese col-
laboration working at the hoary Kolar Gold Mines of the Dharwar Craton in
Karnataka. They observed 6 presumed proton decay events with a half life of
5 × 1030 years. However the above claim turned out to be an illusory mirage,
similar to the pentaquark case of 2002 which was discussed previously. In 1985
the Irvine-Michigan-Brookhaven group in the Ohio salt-mine determined the
proton decay lifetime to be greater than 2 × 1032 years. As this number far
exceeds the predicted age, the SU (5) GUT was ruled out [55].
Returning to the GSW model, we now know that the initial claim that
there is no electric charge quantization in the GSW model was actually wrong.
This is evident from the fact that, if the electric charge is defined as in Equa-
tion 10.7, then any Englert-Brout field with a hypercharge of Yφ does in ac-
tuality possess proper positive and zero charges in Equation 10.9, as well as
all the properly quantised charges defined in Equation 10.8. However, it is to
be noted that the above Equation 10.7 is phenomenological. Below we shall
demonstrate that it is in fact rigorously true, with no arbitrariness whatsoever.
Now the best empirically confirmed theoretical framework of particles to-
day is termed the Standard Model (SM), which is a combination of the GSW
model and the QCD built around the group SU (3)c ⊗ SU (2)L ⊗ U (1)YW . As
per the SM there are three generations of particles with quarks and leptons in
pairs. The first generation (u, d) , (νe , e); the second generation (c, s), (νµ , µ);
and the third generation (t, b), (vτ , τ ). As per the SM these three are basically
repetitive in nature. This empirical reality is theoretically confirmed by the
fact that anomalies (as we shall see below) cancel each other out for each
generation separately. Let us therefore analyze the first generation fermionic
matter particles in detail.
The first generation fermions are assigned to the following representations
for the group SU (3)c ⊗SU (2)L ⊗U (1)YW . (Note that we are taking an arbitrary
number of colour Nc colour groups for reasons which will become evident
below. However one may place Nc = 3 for the empirically confirmed QCD at
any stage).
 
u
qL = , (Nc , 2, Yq )
d L
uR , (Nc , 1, Yu )
dR , (Nc , 1, Yd )
 
νe
lR = , (1, 2, Yl )
e L
eR , (1, 1, Ye ) (10.11)
There are five unknown hypercharges above. Including the unknown Yφ of
376 Group Theory in Particle, Nuclear, and Hadron Physics

the Englert-Brout doublet φ (eqn. (9)), there exist six unknown hypercharges
for the first generation of particles in the GSW model. The question we ask is:
are there enough constraints within the mathematical structure of the GSW
model that these unknown hypercharges are sufficiently constrained to ensure
charge quantisation in the GSW model? For the group SU (2)L ⊗ U (1)YW we
define the electric charge operator in the most general way in terms of the
diagonal generators of the groups as,

Q0 = a0 T3 + b0 Y (10.12)
where T3 and Y are the diagonal generators of SU (2)L and U (1)YW , respec-
tively. We can always rescale the electric charge once as Q = Q0 /a0 and hence
(b = b0 /a0 )

Q = T3 + b Y (10.13)
In the GSW model the SU (2)L ⊗U (1)YW symmetry provides three massless
generators W1 , W2 , W3 of SU (2)L and X of U (1)Y . To provide mass to the
gauge particles for the weak interaction the symmetry was broken using an
Englert-Brout doublet φ in eqn. (9) by Salam and Weinberg. As per their
analysis this provided mass to the W ± and Z 0 gauge particles while ensuring
zero mass for photons γ. Thus U (1)em is the exact consequent symmetry in
the process SU (2)L ⊗U (1)YW → U (1)em . Now so far our discussions regarding
the electric charge have pertained only to the groups SU (2)L ⊗ U (1)YW and
U (1)em . The definition of the electric charge does not involve the colour degree
of freedom. But we know that in each generation, as the quarks partake in
the electroweak interaction of SU (2)L ⊗ U (1)YW as well, and as they are also
aware of the colour degree of freedom, we cannot ignore colour entirely. We
should take account of the colour for quarks, as the anomalies which take
into consideration all the fermions present in the theory will automatically
demand inclusion of the colour degree of freedom in a consistent study of the
GSW model built around the group SU (2)L ⊗ U (1)YW . Hence we are forced
to resort to the larger group SU (3)c ⊗ SU (2)L ⊗ U (1)YW , which is the group
of the SM anyway.
Let us choose the EB doublet in Equation 10.9 as per the standard mecha-
nism [46], [45], [82]. Let the T3 = − 21 corresponding to the Englert-Brout field
develop a nonzero vacuum expectation value < φ >0 . As per the EB mech-
anism for SSB, to ensure that one of the four generators (W1 , W2 , W3 , X) is
thereby left unbroken (meaning that what we are left with is a massless photon
as a generator of the U (1)em group), we demand:

Q < φ >0 = 0 (10.14)


For the Q operator we use Equation 10.13,

T3φ + b < φ0 >= 0 (10.15)


Glashow-Salam-Weinberg Model 377

This fixes the unknown b and we obtain:


1
Q = T3 + ( )Y (10.16)
2Yφ
Note that Yφ exists in the denominator of the expression for the electric
charge.
Anomalies play a very significant role in quantum field theories (QFT)
[46]. In fact, the presence of such anomalies in QFT renders the theory non-
renormalizable. Hence one has to ensure that these anomalies for all the de-
grees of freedom in a particular theory cancel each other out. As we require
the GSW theory to be renormalizable, we have to ensure that all the anoma-
lies theoretically neutralise each other for all the particles. So quantum field
theory demands that anomalies should cancel for all the particle known today:

{(u, d, s, c, b, t), (e, µ, τ, νe , νµ , ντ )}


With all the degrees of freedom known for all particles mentioned, the
above does hold true. But is it possible that anomalies may cancel out for some
subsets of these? Phenomenologically we know that there are three generations
of elementary particles:

{(u, d), (νe , e)} , {(c, s), (νµ , µ)} , {(t, b), (ντ , τ )} (10.17)
This separation of all the particles into three distinct generations is a priori
an experimental fact, implying that these three structures are repetitive in
nature. Astonishingly, it turns out that the anomalies actually neutralise each
other for each generation separately. So as far as anomalies are concerned, one
generation is not aware of the existence of the other. This generation-wise
cancellation of anomalies has deep implications for theoretical models.
Now it is well known in quantum field theory that anomalies are intrinsi-
cally related to topology. Hence one may speculate that due to some – as of yet
unknown – topological reasons the topological anomalies should cancel out,
generation by generation. It may very well be related to the fact that there are
three space dimensions (x,y,z) in SO(3). Viewed this way, separate generation-
wise cancellation of anomalies may be a theoretical constraint which is being
confirmed by experiment.
Now separate generation-wise cancellation of anomalies brings in the re-
quirement for the satisfaction of the following three constraints [31]:

(a) T rY [SU (NC )]2 = 0 (10.18)


which yields

2Yq = Yu + Yd (10.19)

(b) T rY [SU (2)L ]2 = 0 (10.20)


378 Group Theory in Particle, Nuclear, and Hadron Physics

which gives

22 Yl + Nc [22 Yq ] = 0 (10.21)
and thus
Yl
Yq = − (10.22)
Nc

(c) T r[Y 3 ] = 0 (10.23)


giving

2Nc Yq3 − Nc Yu3 − Nc Yd3 + 2Yl3 − Ye3 = 0 (10.24)


To simplify Equation 10.24 we need to know simplifying terms for
Yu , Yd , Ye , in addition to Yq in Equation 10.22.
Now note that the SSB via the Englert-Brout field which generates mass-
less photons, yields an electric charge expression as in Equation 10.16. This
in itself has not generated a non-chiral charge. The charge is still chiral. Note
that it just produces left-handed charges for the left-handed fields and right-
handed charges for the right-handed fields. Hence the SSB does not alter the
chirality aspect of the electric charge itself. This is the point we discussed
above.
Now how do we ensure that we have a non-chiral U (1)em field theory? So
far with the SSB for the Englert-Brout field and the additional requirement
of the cancellation of anomalies, we have determined separate left-handed and
right-handed charges. Thus these are still chiral in nature. To ensure non-chiral
charges of the U (1)em , something more has to be accomplished. And that is
to demand that the above photon must act non-chirally, not distinguishing
between left-handed and right-handed quarks and leptons. So what this new
requirement enforces is that the charges themselves should be immune to the
characteristic of handedness and hence we demand that the electric charges
in U (1)em should satisfy the following property:

QL = QR (10.25)
Imposing this condition in Equation 10.16 we get,

Q(uL ) = Q(uR ) : Yu = Yq + Yφ (10.26)

Q(dL ) = Q(dR ) : Yd = Yq − Yφ (10.27)

Q(eL ) = Q(eR ) : Ye = Yl − Yφ (10.28)


Note that the first anomaly constraint in Equation 10.19 is automatically
Glashow-Salam-Weinberg Model 379

satisfied by these conditions. Now substituting Yq from Equation 10.22 and


Yu ,Yd ,Ye from above into Equation 10.24 one obtains:
3
(Yl + Yφ ) = 0 (10.29)
The equation is reduced to two unknowns Yl and Yφ which contract to

Yl = −Yφ (10.30)
and putting this in Equation 10.22

Yq = (10.31)
Nc
These placed into Equation 10.26 yield,
1
Yu = Yφ ( + 1) (10.32)
Nc
And similarly for Yd and Ye . Finally, substituting these expressions into
Equation 10.16 (and using Q(u) = Q(uL ), Q(d) = Q(dL )),
1 1
Q(u) = (1 + )
2 Nc
1 1
Q(d) = (−1 + )
2 Nc

Q(νe ) = 0

Q(e) = −1 (10.33)
For Nc = 3 these yield the correct charges for the u- and the d-quarks. And
thus we see that the structure of the GSW Model with QCD (as required by
anomaly constraints) gives the correct electric charges for all the particles in
the first generation. Note that inspite of the fact that U (1)em does not know
of colour, the electric charges are actually dependent upon colour itself. Also
note that this charge quantisation is independent of the Englert-Brout field
hypercharge Yφ [89].
Note the fact that the electric charge of the quark has colour dependence
built into itself is a significant new result [34] for the Standard Model. How-
ever this is in direct conflict with what others have been believing and using
throughout. Electric charges of quarks built up by hand as in Equations 10.2
and 10.6 in the early period of the GSW model by their very definition had
no in-built colour dependence. Thus these charges were always Q(u) = 32 and
Q(d) = − 13 . This is well known [82],[45],[46],[54]. In fact in Chapter 6 we
showed that for the Standard Model (constructed as GSW plus QCD) the
above colour dependence is the correct one, while the static charges Q(u) = 23
and Q(d) = − 31 are the wrong charges to use for arbitrary Nc .
380 Group Theory in Particle, Nuclear, and Hadron Physics

For Nc = 3 obviously these charges match, but the static charge does not
know of colour while the other (the correct one) says that it has colour in its
very core. This point also goes against using the incorrect Equations 10.2 and
10.6 for defining the electric charge in the GSW model.

10.4 Millicharged Neutrinos


Another pitfall in selecting Yφ = 1 is to assign a deeper significance to this
value. So there have been papers where people ask why should we take Yφ = 1
exactly and that perhaps it may be a little way off to allow a value such as
Yφ = 1 + δ where δ  1. Now for T = 12 and arbitrary Yφ for the φ-field one
finds that the charge of proton possesses a value Q(p) = 2Q(u) + Q(d), for the
neutron it is Q(n) = Q(u) + 2Q(d) and for the lepton with this new Y φ it is:

1 3 3 (δ/Y φ)
Q(p) = (1 + )+
2 Nc 2 Nc
1 3 3 δ/Y φ
Q(n) = (−1 + )+
2 Nc 2 Nc

Q(e) = −1 − δ/(2Y φ); Q(νe ) = −δ/(2Y φ) (10.34)

————————————————–
Problem 10.1: Using the above method of proving charge quantisation in
the Standard Model with Yφ = 1+δ where δ  1, obtain the above expressions
for charges of the proton and neutron.
————————————————–

Thus using the arbitrary value Yφ = 1 + δ where δ  1 the neutrino charge


is − (δ/2). This is a very small negative value. In other words, the neutrino has
a millicharge. Thus this model predicts the possible presence of millicharged
particles, and indeed much experimental effort has been expended in search
for such entities. In fact there are claims that such millicharged particles may
have been found in the laboratory [90].
However as per our calculation, the Englert-Brout field hypercharge Yφ is
unconstrained in the whole process of charge quantisation. It is not restricted
to any specific value and thus φ → φ + δ does not make any sense. Therefore a
unique prediction of our model is that millicharged particles should not exist
[35],[90].
Glashow-Salam-Weinberg Model 381

10.5 Second Generation Fermions


For the second generation the fermions may be classified with respect to
the group SU (Nc ) ⊗ SU (2)L ⊗ U (1)YW as follows:
 
0 c
; Nc , 2, Yq0

qL =
s L
cR ; (Nc , 1, Yc0 )
sR ; (Nc , 1, Ys0 )
 
0 νµ
lL = ; (1, 2, Y00 )
µ L
µR ; (1, 1, Ye0 ) (10.35)
Now for the second generation fermions, all these hypercharges are a priori
different from the corresponding ones for the first generation hypercharges in
equation (11). Note that in all the present representations by other researchers
(for e.g. [31],[46],[6],[47]) Yφ = 1, Yφ0 = 21 is arbitrarily taken. Thus for those
frameworks the corresponding hypercharges in different generations are always
equal to each other. As we have shown above, without some physical and
mathematical justifications, this is not acceptable.
Now for this generation also independently, the same Englert-Brout field
with T3φ and Yφ quantum numbers would yield the proper charge quantisation.
We obtain independent charge quantisation for the third generation also. Thus
one finds that the electric charge is quantized for the three generations as
follows:
 
1 1
Q(u) = Q(c) = Q(t) = +1
2 Nc
 
1 1
Q(d) = Q(s) = Q(b) = −1 (10.36)
2 Nc
Hence it is evident that charge quantization is an intrinsic property of the
Standard Model – in opposition to earlier wrongly held views to the contrary.
And this holds for an Englert-Brout field with an arbitrary and unconstrained
weak hypercharge Yφ . With all this it was possible to ensure the non-chiral
character of U (1)em . As we shall see below this also provides masses to the
W ± , Z 0 bosons.
However, the fermions are still massless. We can generate masses for
fermions with the Englert-Brout field of a weak isospin doublet. This is be-
cause a φ with a weak isospin doublet has its gauge invariant mass term as
φq¯L qR . These are Yukawa mass terms in the GSW model.
382 Group Theory in Particle, Nuclear, and Hadron Physics

In the GSW model, the Yukawa mass term with T = 12 for φ (where
T3φ = − 12 component develops a non-zero expectation value < φ >0 ) [82]:

L = −φ† q̄L ūR + φqL d¯R + φeL ēR (10.37)


which on demand of gauge invariance yields,

Yu = Yq + Yφ
Yd = Yq − Yφ (10.38)
Ye = Ye − Yφ
Note that these are the same terms as obtained with the QL = QR con-
straint for T3φ = − 12 as in Equations 10.26, 10.27, and 10.28. So for T = 12
field QL = QR constraints provide the same hypercharge relationship as the
Yukawa mass term. Note that the very existence of the Yukawa mass term re-
quires charges to preexist whence QL = QR and corresponding hypercharges
arise before Equation 10.38. So the two are intrinsically related.
This seems to indicate some deep connection between the electric charge
and the mass in U (1)em Quantum Electrodynamics. Now it is known that the
QED theory is a renormalisable theory [46].
QED is renormalizable in terms of two parameters, the bare mass m0 and
the bare charge e0 . When we calculate the true mass m and the true charge
e (the actually measured quantities), we find that these expressions diverge.
Next we regularise the theory by introducing an unphysical regulator Λ such
that the resulting charge and mass are finite. That is

m = m(m0 , e0 , Λ)
e = e(m0 , e0 , Λ) (10.39)
To make sense out of the regularized theory we must somehow make m
and e independent of Λ. Thus we ensure that m0 and e0 are functions of Λ
such that m and e do not depend upon Λ. That is,
dm ∂m ∂m0 ∂m ∂e0 ∂m
=0= + +
dΛ ∂m0 ∂Λ ∂e0 ∂Λ ∂Λ
de ∂e ∂m0 ∂e ∂e0 ∂e
=0= + + (10.40)
dΛ ∂m0 ∂Λ ∂e0 ∂Λ ∂Λ
These coupled equations are of first order and their initial conditions are
provided by the experimentally measured or renormalized mass and charge.
Thereupon one uses the standard renormalization techniques [2]. The point
that should be kept in mind is that the bare mass m0 and the bare charge e0
are two independent quantities.
Note that e0 and m0 may be directly identified with the electric charge
and mass as we have obtained above in the Standard Model. Thus we may
Glashow-Salam-Weinberg Model 383

surmise that the mass of the electron (and also other fermions) is entirely
electromagnetic in origin.
Thus we should distinguish between the Old Standard Model (OSM) which
uses non-colour dependent charges as in Equations 10.2 and 10.6; and the New
Standard Model (NSM) which uses colour dependent charges as in Equation
10.36. In NSM both e0 and m0 arise as independent entities in the same SSB
mechanism. However, this is not true of the OSM. Thus NSM is consistent as
to renormalization (see vixra.org/abs/1603.0344) while OSM is not.

10.6 Right-Handed Neutrino


Now in the above discussion we have ignored the right-handed neutrino
νR in the SM. In recent years, primarily because neutrino oscillations have
become an experimentally confirmed fact, the presence of νR seems to be
required.
So let us include νR in our formalism of the SM as a representation of the
group SU (3)c ⊗ SU (2)L ⊗ U (1)YW . Hence to the first generation of particles,
we have to add another new particle field:

νR : (1, 1, Yν ) (10.41)
Then as in Equations 10.26, 10.27, and 10.28,

Q (νL ) = Q (νR ) : Yν = Yl + Yφ (10.42)


 3
Now the Tr Y = 0 condition with an extra term added to Equation
10.24 becomes:

2Nc Yq3 − Nc Yu3 − Nc Yd3 + 2Yl3 − Ye3 − Yν3 = 0



(10.43)
We already saw in Equation 10.27 that the curly bracket expression is in
fact provided by Equation 10.29. Thus the above equation is actually,
3
(Yl + Yφ ) − Yν3 = 0 (10.44)
Given Yν in Equation 10.42, this equation is automatically satisfied and
thus Tr Y 3 = 0 - with νR included - does not bring in any new condition
on the hypercharges. And so the property of charge quantisation as derived
above (without νR ) is lost when νR in included.
But the above expressions now demand that Q (νL ) = Q (νR ) 6= 0. Given
the fact that there exist a large number of neutrinos in the universe ∼ 1089
(to only ∼ 1080 of baryons), even a small neutrino charge would have very
observable electromagnetic consequences in the universe indeed. Not having
384 Group Theory in Particle, Nuclear, and Hadron Physics

observed any such anomalous effects, we conclude that the charge of neutrinos
is zero. Hence,

Q (νR ) = =0 (10.45)
2Yφ
Note that in the above Q (νR ) = 0 implies that it is YYφν = 0. This means
that either Yν = 0 or Yφ = ∞. Now the latter is ruled out as we saw that
all the electric charges above had factors like YYφ , with Y’s from different
representations, and these will get messed up with this value of Yφ . Hence we
necessarily obtain Yν = 0. Now given the fact that Yν = 0 in the Standard
Model [82], the neutrinos are characterized by an absence of gauge couplings
(as we see below). Hence these are termed “sterile”.
Next, note that for the left-handed charge of the u-quark (and similarly
for the d-quark) we had obtained, as clear from Equations 10.33 and 10.31,
that:
1 1
Q(uL ) = (1 + )
2 Nc
1 Yq
= (1 + ) (10.46)
2 Yφ
Note that N1c is the baryon number . We suggest that this is a general
property and that we should define the baryon number in the SM as:
Yq
=B (10.47)

Here in the SM the baryon number arises as the ratio of the hypercharge
of the left handed quark with respect to the Higgs hypercharge. Since the
Englert-Brout field provides the ubiquitous background uniform structure
within which the particles exist, this is a consistent definition of the baryon
charge for the left-handed quarks.
It is important to note that this so-called global quantum number is chi-
ral in nature. This is consistent, as the weak interaction is only left-handed
anyway.
Similarly for the left-handed electron the electric charge is,
1 Yl
Q(eL ) = (−1 + ) (10.48)
2 Yφ
We now associate a lepton number with
Yl
= −L (10.49)

which yields the correct charges. Notice that this is a natural definition of the
lepton number. Just as in case of the baryon number in the SM, the lepton
Glashow-Salam-Weinberg Model 385

number arises as the ratio of the hypercharge of the left handed lepton with
respect to the Englert-Brout field hypercharge. As such it is natural to treat
this as the lepton number. More so, as the Englert-Brout field hypercharge
remains unconstrained by the theory and the lepton number is thus fixed by
the background Englert-Brout field. And as for the case of the baryon number,
the lepton number is likewise only chiral in nature.
Now we look at the important result for the hypercharge of the right-
handed neutrino,

Yν = 0 (10.50)
What is the significance of this result? We found earlier that it is the ratio
of hypercharges to that of the Higgs hypercharge through which the baryon
number, the lepton number, and the electric charges get defined in the SM.
Thus Yν = 0 is special.
Hence it is immediately evident that one cannot define any lepton number
for this νR . So though the left-handed neutrino exists and is identified by its
lepton number which places it in the left-handed doublet with the electron, the
so-called right-handed neutrino is completely unlike it, and has no associated
lepton number. This νR is colourless, massless, chargeless, weak-isospin-less,
hypercharge-less and lepton-number-less.
Once it is realized that νR possesses no lepton number, it is evident that
there hence cannot be any Dirac mass (Yukawa coupling mass) or Majorana
mass for the left-handed neutrino. So if, as per neutrino oscillation, the neu-
trino indeed has a mass, then we have to be able to generate it in some other
manner. Below we shall demonstrate that this is indeed possible.
To understand it we require Wigner’s analysis [46], [91] of the irreducible
representation of massless entities for the Poincare group (Appendix-F).
For massless fields, he showed that, in addition to the conservation of par-
ity (as for the photon), there exist two states of polarization +h and −h,
where h stands for the helicity. But for massless entities, when parity is not
conserved (as in the case of weak interactions), the two states +h and −h
are actually two independent and different irreducible representations of the
Poincare group. The left-handed neutrino gets lumped with the left-handed
electron by virtue of possessing a lepton number. And therefore the other en-
tity, the so-called right-handed neutrino, being of a different representation,
is actually another independent entity altogether. And this is exactly what
we have determined here. Thus our work is a confirmation of Wigner’s anal-
ysis of Poincare groups. Clearly, the so-called right-handed neutrino has been
completely misunderstood so far.
So νR having a definite helicity decouples from νL first and as a colourless,
massless, chargeless, weak-isospin-less, hypercharge-less and lepton-number-
less entity, is completely inert and looks tantalisingly like a new æther.
386 Group Theory in Particle, Nuclear, and Hadron Physics

10.7 Gauge Bosons and Neutral Currents


Given a global non-Abelian (and Abelian) group, to obtain a local gauge
transformation we invoke a covariant derivative Dµ . So we make the following
replacement for the GSW group SU (2)L ⊗ U (1)Y ,

~ µ − ig1 1 Y Bµ
∂µ → Dµ = ∂µ − ig2 T~ · W (10.51)
2Yφ
where W~ are the three generators of the SU (2)L group and Bµ of U (1)Y
group. Acting on the first generation particles (Equation 10.11) it gives
   
~ ~ Yq
Dµ qL = ∂µ − ig2 T · W − ig1 Bµ qL
2Yφ
   
Yu
Dµ uR = ∂µ − ig1 Bµ uR
2Yφ
   
Yd
Dµ dR = ∂µ − ig1 Bµ dR
2Yφ
   
~ ~ Yl
Dµ lL = ∂µ − ig2 T · W − ig1 Bµ lL
2Yφ
   
Ye
Dµ eR = ∂µ − ig1 Bµ e R (10.52)
2Yφ
Now we write the GSW Lagrangian as:
X
L= f¯ i γ µ ∂µ f (10.53)
(f =lL ,eR qL ,uR ,dR )

For the U(1) group as applied to the lepton this is (subscript f denotes a
leptonic fermion),

   
¯ µ Yl µ Ye
Lf (U (1), lepton) = lL iγ −ig1 Bµ L + ēR iγ −ig1 B µ lR
2Yφ 2Yφ
g1
=+ [Yl (ν̄L γ µ νL + ēL γ µ eL ) + Ye ēR γ µ eR ] Bµ (10.54)
2Yφ
Given

W ± = (−W1 ± iW2 ) / 2 (10.55)
h i
Lf (SU (2), lepton) = −¯lL iγ µ ig2 T~ · W
~ µ lL
Glashow-Salam-Weinberg Model 387

g2 h √ √ i
=+ ν̄L γ µ νL Wµ0 − 2ν̄L γ µ eL Wµ+ − 2ēL γ µ νL Wµ− − ēL γ µ eL Wµ0
2
(10.56)
Note that in U (1)en we have a photon field Aµ (to be defined below in
terms of Wµ0 and Bµ ). Hence the Lagrangian is (note Q is the electric charge)

Lem = QAµ [eL γ µ eL + eR γ µ eR ] (10.57)


Assume that the neutrino (which does not couple to the photon), however,
does couple to the orthogonal neutral Zµ field. From the above it is evident
that:
g1 g2
Zµ ∼ − Yl Bµ + Wµ0 (10.58)
2Yφ 2
Using Equation 10.30 from charge quantization above, Yl = Y φ
g1 g2
Zµ ∼ Bµ + Wµ0
2 2
Orthogonal to it, we have:
g2 g1
Aµ ∼ Bµ − Wµ0
2 2
Properly normalised these are

g1 Bµ + g2 Wµ0
Zµ = p (10.59)
g12 + g22
and

g2 Bµ − g1 Wµ0
Aµ = p (10.60)
g12 + g22
Inverting we get
g2 Aµ − g1 Zµ
Bµ = p (10.61)
g12 + g22
g1 Aµ + g2 Zµ
Wµ0 = p (10.62)
g12 + g22
Looking at the electron terms in the Lagrangian,

g1 g2 g1
(Yl ēL γ µ eL Bµ ) + −ēL γ µ eL Wµ0 + (Ye ēR γ µ eR Bµ )

Le =
2Y φ 2 2Yφ
!
g1 g2 A µ − g1 Z µ
− (ēL γ µ eL ) p
2 (g12 + g22 )
388 Group Theory in Particle, Nuclear, and Hadron Physics
!
g2 g1 A µ + g2 Z µ
− (ēL γ µ eL ) p
2 (g12 + g22 )
!
µ g2 Aµ − g1 Zµ
+ (eR γ eR ) g1 p 2
(g1 + g22 )

(   )
−g1 g2 g1 g2 1 g1 g2
= Aµ ēL γ µ eL − p − ēR γ µ eR p 2
2 2 (g12 + g22 g1 + g12

( ! !)
g 2 − g22 g12
+Zµ ēL γ µ eL p1 + ēR γ µ eR p (10.63)
2 g12 + g12 g12 + g12

As the electromagnetic term is in Equation 10.57 for q = −e, from the


above we obtain:
g1 g2
e= p (10.64)
g12 + g12
writing,
g1 g2
sin θω = p 2 2
; cos θω = p 2 (10.65)
g1 + g2 g1 + g22
and hence,
e e
g1 = ; g1 = (10.66)
sin θω cos θω
The above equations are all well known in the GSW model [46], [45]. Here
these have been obtained through consistency arguments and are independent
of the Englert-Brout field Yφ . Hence the entire structure of the GSW Model
stands intact, while charge quantisation and the entire mathematical edifice is
independent of the EB field hypercharge Yφ , as it is not constrained or pinned
down at all.

————————————————–
Problem 10.2: Solve for the quark sector of the above Lagrangian and
obtain the expressions in terms of Aµ and Zµ coupling as above.
————————————————–
Glashow-Salam-Weinberg Model 389

10.8 Neutrino Oscillation


So far we have assumed the neutrino to be massless. However, there is in
fact no fundamental principle which mandates its masslessness (as for exam-
ple, the masslessness of photon in the U (1)em gauge theory). Through beta
decay in the laboratory, we obtain the upper limit for the neutrino masses as,

mνe < 5.1 ev (mνe  me )


mνµ < 170 kev mνµ  mµ

mντ < 18.2 M eV (mντ  mτ )


While zero mass is consistent with these numbers, yet a small neutrino
mass is still not ruled out due to these limits.
There exist three-flavours of neutrinos - νe , νµ , ντ . If neutrinos are mass-
less then these flavour eigenstates are also mass eigenstates. If one of them
has a (small) mass then it may be that their mass eigenstates denoted by
νi (i = 1, 2, 3) are different from the flavour eigenstates νe , νµ , ντ . Thus one
may obtain neutrino oscillation. This phenomenon of neutrino oscillation
would allow a measure of a (small) neutrino mass. This was first predicted
in 1957, but has been empirically confirmed only in recent years. In fact, the
discoverers of the phenomenon of neutrino oscillation were awarded the Nobel
Physics Prize of 2015.
For the sake of simplicity, let us study the two-flavour case. A priori, we
may take any of the pairs - (νe , νµ ) , (νe , ντ ) , (νµ , ντ ). To be able to compare
to the latest results from the OPERA collaboration, we use the last pair
from above. This flavour case can be transformed to mass eigenstates (ν2 , ν3 ),
whence,
   
|v2 > |νµ >
ψm = =V = V ψF (10.67)
|v3 > |ντ >
is the relationship between the mass eigenstates and the flavour eigenstates
while V is a 2 × 2 matrix.
The conservation of probabilities requires
† †
(ψ m ) ψ m = ψ F ψF ⇒ V † V = 1 (10.68)
demonstrating that V is unitary. We can absorb the complex phases in the
neutrino wave function while making V orthogonal. This is accomplished be-
low.
Inverting the above and re-writing it in terms of the flavour eigenstates
with an orthogonal transformation,
390 Group Theory in Particle, Nuclear, and Hadron Physics

    
νµ cos θ sin θ ν2
= (10.69)
ντ − sin θ cos θ ν3
The time evolution of the flavour eigenstates is defined by:

|νµ >t = cos θ|ν2 >t + sin θ|ν3 >t


= cos θ e−iE2 t |να > + sin θ e−iE3 t |ν3 > (10.70)
and

|ντ >t = − sin θ|ν2 >t + cos θ|ν3 >t


= − sin θ e−iE2 t |ν2 > + cos θ e−iE3 t |ν3 > (10.71)
Where we use non-relativistic quantum mechanics as in Equation 10.67,
ψ m are states of definite mass-energy and thus:

∂ψ m
 
E2 0
i~ = Hψ m = ψm (10.72)
∂t 0 E3
For an initial time (t=0) a pure νµ beam oscillates to ντ . The amplitude
after a time t is given as:
 
< ντ |νµ >t = (− sin θ < ν2 | + cos θ < ν3 |)) cos θe−iE2 t |ν2 > + sin θe−iE3 t |ν3 >

= sin θ cos θ e−iE2 t − e−iE3 t



(10.73)
Thus the probability of obtaining a ντ at some later time (t > 0) from an
initial (t = 0) pure νµ beam is:

 
1
Pνµ → ντ (t) = | < ντ |νµ >t |2 = sin2 (2θ) sin2 (E2 − E3 ) t (10.74)
2

From this the probability of an initial pure νµ still remaining a pure νµ


after a time t is given by,

Pνµ →νµ (t) = | < νµ |νµ >t |2 = cos 2 θe−iE2 t + sin2 θe−iE2 t |2


 
1
= 1 − sin2 (2θ) sin2 (E2 − E3 ) t
2
whence

Pνµ →νµ = 1 − Pνµ →ντ (10.75)


Now the above energies E2 and E3 are non-relativistic in nature, on ac-
count of their derivation from a non-relativistic Schroedinger equation. The
Glashow-Salam-Weinberg Model 391

next step undertaken by conventional investigators ([45] p.487) is, however,


of a very questionable and, indeed, Janus-faced, nature. As Mann observes,
“... make an approximation that goes beyond what I have assumed in using
the non-relativistic Schroedinger equation”. This ‘approximation’ is in fact a
relativistic calculation, which the conventionalist approach resorts to in direct
contradistinction to the immediately preceding non-relativistic derivations.
After utilising the non-relativistic Schroedinger equation, a sudden jump is
made to the relativistic approach.
Now using the relativistic energy for neutrinos,
s
(m2 c4 )
q
E2 = p2 c2 + m22 c4 = pc 1 + 22 2
(p c )
m2 c4
 
∼ pc 1 + 22 2 (10.76)
p c
Assuming that all the high energy neutrinos possess the same momenta
(this is also a major assumption),

m2 c4 m2 c4
   
E2 − E3 = pc + 2 − pc + 3
2pc 2pc
c4
m22 − m23

=
2pc

c4 ∆ m2 c4
= ∆(m2 ) ≈ (10.77)
2pc 2E
where E is the average energy of either of the neutrinos. In time t the path
travelled is Lc and so,
2 !
2 2 |∆ mc2 |L
Pνµ →ντ (t) = sin (2θ) sin (10.78)
4E~c
and Pνµ →ντ (t) is obtained by using Equation 10.75.
Such an analysis can easily be generalised to three flavours. And similar
oscillations are predicted. 2
Note that in Equation 10.78 we have written ∆ mc2 , i.e., the absolute

value of the mass-squared difference. Note that a few important points arise:
2
(1). Why do we need the mass-squared term mc2 in the oscillation?
Why did we not get the first power of mass here? Clearly this is due to the
mathematical structure that has gone into theory. The first basic mathemat-
ical structure is the non-relativistic Schroedinger equation. Now this in itself
may have permitted the first power of mass terms. However the second im-
portant mathematical structure demanded was a contradictory requirement,
namely that of using the relativistic equation for the energy of neutrinos,
namely as E 2 = p2 c2 + m2 c4 . In this formalism the mass-squared terms arise
392 Group Theory in Particle, Nuclear, and Hadron Physics

fundamentally. So the problem is how to reconcile these contradictory condi-


tions of using the non-relativistic Hamiltonian to obtain the energies E2 , E3
and thereafter applying the relativistic energy expressions to implement the
m2  p condition. We have to resolve this issue before we can accept the
above probability expressions as being valid to obtain masses from neutrino
oscillations. Below we show that this can be done consistently, provided we
take the neutrinos to be superluminal.
2
(2). What does ∆ mc2 mean? One answer is that it does not distin-

guish between (E2 − E3 ) and (E3 − E2 ) in Equation 10.78. This means that
whether we start with an initial beam of νµ and study its oscillation to ντ , or
we commence with an initial pure beam of ντ and analyse it oscillation to νµ ,
the result would be the same.
(3). The experimental setup to detect Pvµ → vµ (t), where one starts with
an initial beam of νµ and studies the depleted amount of νµ , is denominated a
disappearance experimental, i.e some νµ have disappeared in this respect
from the initial number.
4. In the appearance experiment, one searches for a new flavour of
neutrino, which was absent in the initial beam and can arise only through
oscillation. Hence Pvµ → vτ (t) corresponds to the appearance of ντ in a pure
initial beam of νµ
Several experiments on the neutrino in the laboratory and on terrestrial
neutrinos have confirmed the phenomenon of neutrino oscillation and hence
indicate a non-zero value of ∆(m2 ) for (ve − vµ ). Thus clearly neutrinos do
have non zero mass.
Here we emphasise the study of the OPERA experiment at CERN. It turns
out that laboratory experiments are best suited to measuring Pνµ →νµ (t) and
Pνµ →νe (t) [45]. These experiments confirmed the neutrino oscillation, but a
more interesting result was the following.
In the OPERA (CERN) and LNGS (Gran SASSO, Italy) experiments,
neutrinos produced at Geneva, CERN were observed at Gran Sasso (Italy), a
distance of 730 km. They were observing ντ in νµ oscillation.
In September 2011 the teams claimed to have observed superluminal neu-
cm
trinos. They asserted the neutrinos exceeded the speed of light c = 3×1010 sec.
5 cm
by 6 × 10 sec. . This occasioned great excitement, because as per the theory
of relativity, nothing can travel faster than light. Naturally, these supposedly
faster-than-light neutrinos were breaking this limit. Later in June 2012, how-
ever, the claim was withdrawn [92]. Hence this appears to have been another
experimental mirage - as in the illusory case of the pentaquark - Chapter 6.
Given that the relativistic energy expression is:

E 2 = p2 c2 + m20 c4 (10.79)
s
m20 c4
E = pc 1 + (10.80)
p 2 c2
Glashow-Salam-Weinberg Model 393

= pvp (10.81)
when
s
m20 c4
vp = c 1 + (10.82)
p2 c2
vp is termed the phase velocity. And the group velocity is given as
∂E c
vg = =q (10.83)
∂p m2 c4
1 + p20c2
and that

vp vg = c2 (10.84)

————————————————–
Problem 10.3: Given the particle velocity v and γ = q 1 with E =
2
1− vc2
m0 γc2 , p = m0 γv, demonstrate that vg = v (i.e. the group velocity is equal
to the particle velocity and that vp > c).
————————————————–

In 1923, Louis de Broglie proposed that matter possesses the same


Manichaean wave-particle duality as light, with a similar relation between fre-
quency and energy. As consequences of time dilation and the relation E = ~ω,
De Broglie concluded that there must be associated with a particle travelling
at speed v, a wave that travels at phase speed vp = q c v2 = βc . Since this
1− c2
speed is greater than c, he referred to this wave as “fictitious”. The conven-
tional answer to this oddity is to note that the group velocity of a wave packet
will be less than c, and to emphasise that, it is this that must be associated
with the velocity of the particle.
Next, suppose the given Lagrangian is defined in terms of a generalised
coordinate q as:

L = L(q, q̇) (10.85)


where q̇ is the time derivative of q. The Lagrange form of the equation of
motion is:
d ∂L ∂L
− =0 (10.86)
dt ∂ q̇ ∂q
One defines generalised momentum as,
394 Group Theory in Particle, Nuclear, and Hadron Physics

∂L
p= (10.87)
∂ q̇
The Hamiltonian of the system is given by

H(q, p) = pq̇ − L (10.88)


This relationship between L and H is denominated the Legendre transfor-
mation in which L is regarded as a function of q and q̇ while H is taken as a
function of q and p. The Hamiltonian form of the equations of motion are:
∂H ∂H
ṗ = − , q̇ = (10.89)
∂q ∂p
Now note that in deriving the above equation for neutrino oscillations, the
mass eigenvalues were eigenstates of the non-relativistic Hamiltonian to pro-
duce the energy in Equation
p 10.72, i.e., H | ν2 >= E2 | ν2 >. However when
we substitute E2 = p2 c2 + m22 c4 in Equation 10.76 (to take account of the
fact that the neutrino rest mass is small compared to the total energy), we are
utilising a relativistic expression. So in one breath we use non-relativistic me-
chanics and in the next breath we are utilise relativistic mechanics to describe
the same entity in the identical situation.
Below we demonstrate that we can reconcile the above contradiction -
provided we take the neutrino to be superluminal, i.e., as travelling with a
velocity greater than that of light.
It is evident from the above dissertation that the relativistic energy can
be written in terms of the phase velocity as E = pvp , with,
s
m 2 c4
vp = c 1 + 20 2 (10.90)
p c
Assume that our neutrinos obey this relationship and move with velocity
vp . As vp > c these are superluminal. Below we see that this assumption
leads to consistency in equations of motion responsible for providing us with
neutrino oscillation.
Now knowing that for a high-energy neutrino, the rest mass energy is much
smaller then its energy:

1 m20 c4
vp ∼ c(1 + )
2 (p2 c2 )
vp = c + vf (10.91)
With

m20 c4
vf = ·c (10.92)
2p2 c2
Glashow-Salam-Weinberg Model 395

vf stands for velocity of neutrino in excess of the velocity of light c and


vf << c

E = pc + pvf
(E − pc) = pvf (10.93)
Now assume that there exists a Lagrangian L(q, vf ) as a function of gen-
eralised co-ordinates q and q̇ = vf . Then as per the Legendre transformation
in Equation 10.88, we define the Hamiltonian as,

H(q, p) = pvf − L(q, vf ) (10.94)


And so,

(E − pc) − L(q, vf ) = pvf − L(q, vf )

= H(q, p) (10.95)
Now for the equations of motion,
∂H
vf = q̇ =
∂p
are good and we require,
∂H ∂L
ṗ = − = =0 (10.96)
∂p ∂q
Now to the virtue of having two first-order equations of motion for the
Hamiltonian vis-a-vis a single second-order equation of motion for the La-
grangian. As we see below, this demands first-order superluminal Hamiltonian
equations of motion and none of the second order. In the above analysis, the
Hamiltonian structure holds even if the Lagrangian is set to zero or is a con-
stant. A motivation for this approach may be as follows. For a non-zero (or
constant) Lagrangian, with the Lagrange equation of motion in the canonical
manner, the linear momentum is conserved. This in turn would imply that
space is homogeneous. Hence if the above conjecture is valid, it would be so
for any homogeneous space - including ours. Thus superluminosity would be
valid for all particles. But this is empirically not so. And hence for Equa-
tion 10.96 to hold, we have to demand that the Lagrangian itself is zero or
a constant. Whence the above equation will be valid but would not demand
homogeneity of space. It will just mean that all the neutrinos will possess the
same and conserved momentum. Remember that we had previously assumed
that all the neutrinos would have the same momenta at high energies (see
statement before Equation 10.77). Now this is the constraint demanded by
the above equation. Thus,

H(q, p) = pvf − L(constant) (10.97)


396 Group Theory in Particle, Nuclear, and Hadron Physics

where L(constant) may be also set equal to zero. Thus we have obtained a
proper Hamiltonian which produces the correct energy difference in Equation
10.77 and yields the above good results for neutrino oscillation in a consistent
manner. What is important is that the neutrino is superluminal and that the
Hamiltonian depends upon vf , the velocity above the velocity of light c. Hence
superluminosity of neutrinos is a clear prediction of our model here.
Several experiments completed so far have obtained subluminal velocities
of neutrinos [92], [93]. However the point to note is that all these experi-
ments have used neutrino disappearance data to obtain the above result. Even
OPERA – which has until the end of 2015 observed only five τ -neutrinos in
an initial beam of µ-neutrinos – still used disappearance data to derive the
neutrino velocity. They need many more tau-neutrinos to be able to obtain
the neutrino velocity in the appearance mode. And this is precisely what our
model predicts. It says that in the appearance mode the neutrinos shall be
seen to be travelling with velocities greater than that of light. These and others
performing similar experiments, or studying Pνe →νµ (t), are urged to confirm
this clear-cut prediction of our model.
In general it is standard to take an imaginary mass to obtain tachyonic
neutrinos. This requires various kinds of model dependencies [94]. However
our result here is more basic and fundamental in nature. In essence, actually
it is “model-independent”. Its inputs are just these: (1) if neutrinos oscillate,
then this phenomenon necessarily requires the non-relativistic Schroedinger
equation and (2) also these neutrinos have very small masses and travel with
high velocities and this necessitates relativistic energies. This is how everyone
derives the above mass squares expression. But these two are in conflict with
each other. What is shown here is that there is no conflict if the neutrino
travels with superluminal velocities.
Now to the issue of a theoretical understanding of the mass of neutrinos.
We have seen that in the GSW model the fermions have masses arising from
the Yukawa coupling gf L φfR where φ is a vacuum expectation value and
both the left-handed and right-handed fields exists. Now, νR ’s existence is
indicated by neutrino experiments. Then assuming that they exist, a priori the
GSW model does permit a neutrino mass term of this kind. It is denominated
the Dirac mass term for neutrino. So the GSW model is entirely capable of
producing Yukawa mass terms for it. The only requirement is that νR , as a
counterpart of νL , must exist in the GSW model. This means that νR - like
νL - has an associated lepton number. As per conventional understanding, the
only difficulty with the Dirac term in the GSW model is to comprehend why
its value is so low, i.e., the issue of why (mv  me ) within the structure of
the GSW model.
One may write another mass term for neutrinos as they are electrically
neutral. It is of the form v cL vR where the RH singlet neutrino couples to the
LH antiparticle. A fermion which has this property of the charge conjugate
being equal to itself is termed a Majorana fermion. Hence the above is a
Majorana mass term.
Glashow-Salam-Weinberg Model 397

Thus a priori within the Standard Model there are two kinds of mass
terms possible for the neutrino – one of the Dirac kind and the other one
of the Majorana kind. At present there are several frameworks which require
only one kind of neutrino, i.e., the Majorana kind (as the Dirac kind above is
not favoured) or some models invoke both of these varieties in some suitable
combination.
If the neutrino mass is of the Majorana kind, then there would be several
experimental signatures of its Majorana nature. One prediction is Majorana
neutrino’s unique signature in the so called, double-beta-decay experiments.
If the neutrino is a Majorana particle it should participate in neutrino-less
double-beta-decay, i.e., 0νββ. However as per latest empirical data this has
been practically ruled out by the Exo-collaboration experiment [95] on Xe138 .
This group observed 22,000 events of double-beta-decay with neutrinos 2νββ,
but none of the kind of neutrino-less double beta-decay 0νββ. This speaks
against the Majorana nature for the generation of neutrino mass.
As discussed above, within a consistent study of the Standard Model,
vR – right-handed neutrinos – possess no associated lepton number. And
thus theoretically both the above mass terms, the GSW Dirac kind, and the
Majorana kind are zero. Hence no first power mass term exists in the GSW
model. But this is fine as it is consistent with the fact that neutrino masses as
extracted in the neutrino oscillation experiments are actually mass-squared.
So the question here is, whether it is possible that a particle possesses zero
mass in the first power, but still has a non-zero squared-mass? Let us explore
the question as to wherefrom the mass-squared term may arise. Assuming
that there is a unitary matrix S connecting the flavour and mass eigenstates
of left-handed neutrinos as
   
νe ν1
νµ  = S ν2  (10.98)
ντ f lavour ν3 mass
Now in the Appendix B we demonstrated that a square complex matrix
can be written as a sum of an Hermitian and an anti-Hermitian matrix,

1  1
A + A† + A − A† = H + K note H † = H, K † = −K (10.99)
 
A=
2 2
where H is Hermitian and K anti-Hermitian. Thus on exponentiation,

exp (arbitary complex matrix) = exp (antiHermitian n × n matrix)

⊗Hemitian n × n matrix (10.100)


= unitary matrix ⊗ Hermitian matrix (10.101)
Hence an arbitrary non-singular complex matrix can always be written as
the product of a unitary matrix U and a Hermitian matrix H,
398 Group Theory in Particle, Nuclear, and Hadron Physics

M = UH (10.102)
Note that,


M † M = (U H) (U H) = H † U † U H = HU −1 U H = H 2 (10.103)

And hence M † M is Hermitian and positive-definite and thus can be diag-


onalized by a unitary matrix S,

S † (M M † )S = Md2 (10.104)
where Md2 are positive and real. Using Equation 10.103 we obtain,

(νmass (S † (M M † )S)νmass ) = Md2 (10.105)


where Md is already diagonal and positive in the mass-squared term. So a sin-
gle unitary transformation on a pure left-handed basis yields a chiral diagonal
mass-square. This is the mass-squared term arising in the neutrino oscillation
expressions. So no Yukawa terms or Majorana first-power mass terms for left-
handed neutrinos, but a mass-squared term is obtained in the GSW model.
Hence, one need not go beyond the Standard Model and the GSW model
which themselves provide the mass-squared term in the neutrino oscillation
expressions.
Note that in the Newton’s Law of Gravitational motion, on fundamental
grounds only, products of two masses arise. For the same body, this may be
taken as a mass-squared self-interaction term. Thus because of the neutrino’s
mass-squared nature, we may surmise that these may act as a gravitational
mass. Thus the GSW provides a mass-squared neutrino which would partic-
ipate as Dark Matter (DM) particles. The weakly interacting massive parti-
cles (WIMPS) such as neutralinos (from super-symmetric theories) and left-
handed neutrinos as in the GSW model (demonstrated here) are viable poten-
tial candidates for being DM particles. Even if neutralinos or super-symmetric
theories are ruled out on experimental grounds (as seems to be happening at
present in early 2016), the new mass-squared neutrino here would still be vi-
able candidates for DM. Note that the word WIMP is just a generic name.
Thus even if neutralinos are ruled out, mass-squared neutrinos (as shown here)
could fulfill the role of WIMPs in our model [96]. And hence all the effects
that we had visualized as affecting the Earth, such as Dark Matter causing
mass extinctions [97] and Dark Matter representing a new source of geother-
mal energy [96], are relevant and valid. Exo-planets in recent years, continue
to confirm the validity of these ideas.
Glashow-Salam-Weinberg Model 399

10.9 Appendix F: Lorentz and Poincare Groups


Homogeneous Lorentz Transformation: Let two events in 4-
dimensional space-time be defined by points (t, x, y, z) and (t + dt, x + dx, y +
dy, z + dz). The ‘distance’ between two points is denominated the interval ds
[46], [91], [41]. We demand that the interval ds be the same for all inertial ob-
servers, i.e. it must be invariant under Lorentz transformations and rotations.
Hence:

ds2 = c2 dt2 − (dx2 + dy 2 + d2z ) (10.106)


With the above notation, time-like separated event intervals have ds2 > 0,
space-like intervals ds2 < 0 and null or light-like intervals ds2 = 0.
In 3-dimensional space dr2 = dx2 + dy 2 + dz 2 is invariant under rotations.
Being a sum of squares, it is positive definite. But in 4-dimensional space-time
the invariant internal is no longer positive definite. Let us define,

contravariant vector : xµ = (x0 , x1 , x2 , x3 ) = (ct, x, y, z)


(10.107)
covariant vector : xµ = (x0 , x1 , x2 , x3 ) = (ct, −x, −y, −z)

and thus
3
X
ds2 = dxµ dxµ = c2 dt2 − dx2 − dy 2 − dz 2 (10.108)
µ=0

Metric tensor gµν helps in transforming from contravariant vector to co-


variant vector (or vice-versa by g µν ) as

X
xµ = gµν xν
ν
 
1 0 0 0 (10.109)
 0 −1 0 0 
with gµν = g µν =
 0 0 −1 0 

0 0 0 −1

Note the above metric tensor (valid in our Minkowski space-time relevant
in Special Relativity), being fixed in advance, plays a passive role. In General
Relativity, however, it depends on matter and even vacuum structure (though
we shall not discuss that here).
We define differential operators
∂ 1 ∂ ~ ~ = ∂ ∂ ∂
∂µ = =( , ∇) {∇ , , }
∂xµ c ∂t ∂x ∂y ∂z
400 Group Theory in Particle, Nuclear, and Hadron Physics

1 ∂ ~
and ∂ µ = g µν ∂ν = ( , −∇)
c ∂t
Lorentz invariant D’Alembertian operator

1 ∂2 ~2
 = ∂ µ ∂µ = −∇
c2 ∂t2
Energy-momentum four vector of a particle
E E
pµ = ( , p~), pµ = ( , −~
p)
c c
with Lorentz invariant
E2
p2 = pµ pν = − p~2 = m2 c2
c2
Generalizing the concept of an orthogonal group O(n), if the group of
transformation of a space with coordinates (x1 , x2 · · · xn , y1 , y2 , ...ym ), leaves
the quadratic form (x21 + · · · x2n ) − (y12 + · · · + ym
2
) invariant, it is termed the
orthogonal group SO(m,n). Thus we see that the Lorentz group is SO(1,3).
(Note the order of (1,3) here, as this is determined by the way we defined our
four vectors xµ (ct, ~x) ).
Four-vector representation plays a basic role in SO(1,3) and is its funda-
mental representation. As products these four vectors produce all the tensorial
representations.
Hence the product of two four-vectors denoted as 4:

4⊗4=1⊕6⊕9

yields a scalar denoted by 1, an antisymmetric tensor by 6 and a traceless


symmetric tensor by 9.
Note that the above analysis is valid for tensorial representations only.
For solely in this case is the four-vector taken as the fundamental representa-
tion of the Lorentz group. When we include the spinorial representation (as
below) then the four-vectors are no longer the fundamental representations
of the Lorentz group. As we shall demonstrate below, then all the represen-
tations of the Lorentz group will be built from the spinorial representations
( 21 , 0) and (0, 12 ).
But before that, let us look at the adjoint representation of SO(1,3). It
has the same dimension as the number of generators of the group. As we saw
in Chapter 3, the adjoint representation is defined in terms of the structure
constants of the corresponding Lie algebra as

(T a )bc = −if abc


The Jacobi identity is actually its Lie algebra in disguise. For the Lorentz
Glashow-Salam-Weinberg Model 401

group thus the adjoint representation has dimension six. It is given by an


antisymmetric tensor Aµν . The adjoint representation plays a basic role in
non-Abelian gauge theories, as we shall see in the Standard Model of particle
physics, built around the group structure SU (3)C ⊗ SU (2)L ⊗ U (1)Y .
Now the Lorentz group SO(1,3) has a transformation which leaves it in-
variant. This is just the rotation group SO(3), generated by the three rotations
in (x, y), (x, z) and (y, z) planes: These are produced by the three rotation
generators (see Chapter 3) supplemented with the t-space,

   
0 0 0 0 0 0 0 0
 0 0 0 0   0 0 0 1 
jx = −i   ; J = −i  ; (10.110)
 0 0 0 1  y  0 0 0 0 
0 0 −1 0 0 −1 0 0
 
0 0 0 0
 0 0 1 0 
Jz = −i 
 0 −1 0 0 
 (10.111)
0 0 0 0
Next we have three more transformations in the (t, x), (t, y) and (t, z)
planes which leave (t2 − x2 ), (t2 − y 2 ) and (t2 − z 2 ) invariant, respectively.
A transformation that leaves (t2 − x2 ) invariant is called a boost along the
x-axis. Note that pure ”boost” Lorentz transformations are those connecting
two inertial frames S and S 0 , moving with relative speed v. These Lorentz
transformations are

0 0 0 0
x0 = γ(x0 + βx1 ), x1 = γ(βx0 + x1 ), x2 = x2 , x3 = x3 (10.112)

where

v2 − 1 v
γ = (1 − 2
) 2 , β = , x0 = ct, x1 = x, x2 = y, x3 = z (10.113)
c c
Here −1 < vc < 1. In the small vc limit the transformations reduce to
the velocity transformations of classical mechanics. Thus the classical veloc-
ity transformation gets generalised at relativistic velocities as in a Lorentz
transformation.
Since −1 < vc < 1 we write vc = tanh φ with −∞ < φ < +∞. Then as
γ − β2γ2 = 1
2

γ = coshφ, γβ = sinhφ

Then the above transformation in terms of variable φ is,


402 Group Theory in Particle, Nuclear, and Hadron Physics

 | 
x0
  0 
coshφ sinhφ 0 0 x
|
x1   sinhφ coshφ 0 0   x1
   
=
  
x2
|
0 0 1 0   x2
 
 
x3
|
0 0 0 1 x3

The variable φ is termed the rapidity.


Note the Lorentz group is defined in terms of six parameters in a differen-
tiable and continuous manner and hence it is a Lie group. However, one pa-
rameter, the boost velocity |~v |, ranges over a non-compact interval 0 6 |v| < 1.
Therefore the Lorentz group is non-compact.
We define the generators K’s of the above boost transformation along the
x − axis (calling the matrix as M):
 
0 1 0 0
1 ∂M  1 0 0 0 
Kx = |φ=0 = −i 
 0 0 0 0 

i ∂φ
0 0 0 0
Similarly the other boost generators
   
0 0 1 0 0 0 0 1
 0 0 0 0   0 0 0 0 
Ky = −i   1 0
 ; kz = −i  
0 0   0 0 0 0 
0 0 0 0 1 0 0 0
This infinitessimal pure Lorentz transformation L is
L = I − i∆φ K ~  v̂ when v̂ = ~v
|~
v|
Exponentiation yields finite pure Lorentz transformations:
~
L = e−iφK·v̂ (10.114)
Thus we see that the most general (homogeneous) Lorentz transformation
has three rotations about the three axes, and three boosts in these directions.
The six generators are as given above.

————————————————–
Unsolved Problem 10.1: Check that the following commutators are
satisfied.
————————————————–

These six generators obey the commutation relations, giving the following
Lie Algebra of the Lorentz group.
Glashow-Salam-Weinberg Model 403

[Ji , Jj ] = iijk Jk
[Ji , Kj ] = iijk Kk
(10.115)
[Ki , Kj ] = −iijk Jk
(i, j, = 1, 2, 3)

The first expression is the Lie Algebra of SO(3) or SU(2) and defines the
angular momentum algebra. The behaviour of the second one is due to the
fact that it can be demonstrated (through some further mathematics) that K ~
is a spatial vector.
Also note that the Lorentz group Lie algebra shows that the pure Lorentz
transformations do not in themselves form a group as the K ~ generators do
not form a closed algebra under commutation.

————————————————–
Problem F.1: Demonstrate that the commutator of two infinitessimal
boosts in different directions (say x-and y-) leads to a rotation about the
z-axis.
————————————————–

Note that the finite homogeneous Lorentz transformation is written as


~ ~
Λ = e−i{φK.v̂+θJ.n̂}

We saw above that the four-vector forms the fundamental representation


of SO(1,3). From this we can form other tensor representations. Note that in
SO(1,3) if we take the time variable t as purely imaginary, then the group
SO(1,3) transforms to the Euclidean version of SO(4). Now we have tensorial
representations akin to those of SO(3).
But as in S0(3) we have tensor representations built from integer spins,
while SU(2), obeying the same Lie algebra as that of SO(3), picks up spin − 12
spinorial representation. Similarly we seek spinors as enlarged representations
of the Euclidean version SO(4) of the Lorentz group SO(1,3).
Let us see how Pauli spinors transform under the Lorentz transformations.
Define

~ = 1 (J~ + iK)
A ~
2 (10.116)
~ = 1 (J~ − iK
B ~
2
Then the commutation Equation 10.115 becomes
404 Group Theory in Particle, Nuclear, and Hadron Physics

[Ai , Aj ] = ijk Ak
[Bi , Bj ] = ijk Bk (10.117)
[Bi , Bj ] = 0 (i, j = 1, 2, 3)

————————————————–
Unsolved Problem 10.2: Confirm the Lie algebra in Equation 10.117
by using Equation 10.116 in the Lie algebra in Equation 10.115.
————————————————–

We have two copies of Lie algebra SU(2) generated by A ~ and B.~ These
are independent of each other as these two SU(2)’s commute. Thus we have
written the Lorentz algebra as SU (2) ⊗ SU (2). Group theoretically SU (2) ⊗
SU (2) is to the group SO(4) what SU (2) is to SO(3).
So as the Lorentz group is essentially SU (2) ⊗ SU (2) and thus states
transforming in a well defined way shall be labelled by (j, j 0 ), i.e., two spin
1/2 angular momenta, the first are corresponding to A and the second one
corresponding to B.
As a special case of (j, j 0 ) let us put either one of these as zero:

(j, 0) →J~ = +iK


~ ~ = 0)
(B
(10.118)
(0, j) →J~ = −iK
~ ~ = 0)
(A

Thus there are two types of spinors as above. In the literature they are
often referred to as L-handed and R-handed spinors.
In general, as we saw above the Lorentz algebra can be labelled by two half
integers (j, j 0 ). The dimension of the representations is (2j + 1)(2j 0 + 1). The
generators of the rotation J~ is related to A ~ and B ~ by A ~ + B.
~ Therefore, by
0
the addition of angular momenta, in the representation (j, j ) we have states
of all spins j in integer steps between the values of |j − j 0 | and j + j 0 .
Note that the rotation group is compact and the homogeneous Lorentz
group is non-compact. The angle of rotation in the rotation group extends
from φ = 0 to φ = 2π and by identifying these points, the line closes into a
circle. Hence the rotation group is compact. While the Lorentz boost vc → 0
and vc → 1 is an open interval and thus it is non-compact. The group space
of the rotation group is finite while that of the Lorentz group is infinite.
From quantum mechanics we know that the unitary representations of non-
compact groups are infinite dimensional. What we saw for Lorentz group was
a finite and non-unitary representation (say L-handed and R-handed spinors)
of Lorentz group.
Wigner pointed out that the basic group defining particle physics is not the
homogeneous Lorentz group as given above, but the inhomogeneous Lorentz
Glashow-Salam-Weinberg Model 405

group, called the Poincare group, which contains in addition to the Lorentz
boosts and rotations, translation in space-time as well.
Poincare Group: Through the spinorial representation we studied spin
as a “kinematic” structure of fundamental entities. But mass through

M 2 = P µ Pµ (10.119)
arises from the momentum operator, not available in homogeneous Lorentz
transformations because it arises as a generator of the space-time translations.
Hence the translation group is given as
µ
e−iP aµ
(10.120)
where aµ is the parameter of translations in xµ → xµ aµ . This trans-
lation added to the (homogeneous) Lorentz transformation is called the
non-homogeneous Lorentz transformation labelled ISO(1, 3) of the Poincare
group. Now we know that translations commute

[pµ , pν ] = 0

The Poincare algebras are defined as

[J i , J j ] = iijk J k , [J i , K j ] = iijk K k , [J i , P j ] = iijk P k

[K i , K j ] = −iijk J k , [P i , P j ] = 0, [K i , P i ] = iP 0 δ ij

[J i , P 0 ] = 0, [P i , P 0 ] = 0, [K i , P 0 ] = iP i

The first row above, K i and P i are vectors under rotations of which J i
are the generators. P 0 is time translation and as J i and P i commute with it,
these are invariants of motion. But K i is not conserved and thus eigenstates
cannot be labelled by eigenvalues of K. ~
At the classical level, a Lagrangian description is found to be useful in
specifying the dynamics of the system. In quantum field theory however we
would like to understand how the concept of a physical particle arises from the
quantised fields. Let the Poincare transformation be represented by a unitary
matrix and the generators J i , K i , P i and P 0 by Hermitian operators.
Let us determine all the Casimir operators which will label or classify the
particle states. The first Casimir operator is

C1 = P µ Pµ (10.121)
406 Group Theory in Particle, Nuclear, and Hadron Physics

One can easily check that C1 commutes with all the generators of the
Poincare group. For a single-particle state it has the value m2 where m is the
particle mass.
How do we now specify the spin of the particle? It is not J~2 , as J~2 does
not commute with all the generators of the Poincare group.

————————————————–
Problem F.2: Demonstrate that J~2 does not commute with all the gen-
erators of the Lorentz group.
————————————————–

We define a new so called Pauli-Lubanski pseudo-vector Wµ :


1
Wµ = − µνρσ J νρ P σ (10.122)
2
where µνρσ is a totally symmetric four-vector; J νρ = −J νρ , i.e., an antisym-
metric Lorentz generator. We shall not go into details here but we can obtain
the second Casimir operator as

C2 = Wµ W µ = −m2 s(s + 1) (10.123)


where s is the spin of the particle.
So for a massive particle m 6= 0, we can label the representation by mass
m and spin s. Note that we may perform a Lorentz transformation to bring
P µ into the form P µ = (m, 0, 0, 0). This provides us the freedom to still
perform the spatial rotations. Therefore the representation with momentum
pµ = (m, 0, 0, 0) allows it be the basis for spatial rotations. Now a subgroup
of Poincare group that leaves pµ invariant is termed the little group. It has
the same structure for all momenta in pµ .
Here as we want to work in a spinor representation, we take the little
group as SU(2). So the massive representation of the Poincare group is given
by mass m and spins s = 0, 21 , 1, 23 , . . . where each state of spin s has extra
labels of s3 = −s, −s + 1, . . . , +s. So a massive particle of spin s has (2s + 1)
number of states.
For P 2 = 0 case the rest frame does not exist. We may choose in the time-
like case the corresponding rest frame as P µ = (k, 0, 0, k). This describes a
massless particle in motion along the z-axis.To obtain the little group in this
case, we have to know what is the most general Lorentz transformation that
will leave P µ invariant. We can see that the rotations in (x,y) plane will leave
P µ invariant. Thus this is an SO(2) group generated by J 3 .
For SO(2) being an Abelian group the irreducible representations are one-
dimensional. For SO(2) rotations in (x, y) plane the generator is the angular
momentum J 3 . So this one-dimensional representation is labelled by the eigen-
value h of J 3 . It represents the projection of the angular momentum in the
direction of motion, in this case the z-axis and is denominated the helicity.
Glashow-Salam-Weinberg Model 407

One requires topological arguments to prove that the helicity h is quantised


as: h = 0, ±1/2, ±1, . . . . Thus the massless particle possesses helicity as the
only quantum number specifying it.
A U(1) rotation of the little group for a state of helicity h is,

U (1) = e−ihφ (10.124)


Thus a massless particle with helicity −h will be a different and indepen-
dent kind of particle from the above +h helicity particle. Thus +h and −h
helicity states would correspond to different representations of the Poincare
group.
Now the helicity projection of the angular momentum along the direction
of motion is:

h = pz Jz /|pz | (10.125)
So under parity transformation h changes sign. Consequently, whether
h will be conserved or not will depend upon the particular interaction the
massless particle partakes in. Thus in electromagnetic interactions parity is
conserved and hence the +h and −h may both describe the same photon states
of polarization. Thus a photon may exist in a right-handed and a left-handed
state.
Gravitational interaction also conserves parity and thus for the massless
graviton only two polarisations states of h = +2 and h = −2 may exist. This
explains as to why the graviton, in spite of having a spin of 2, has only two
states of polarization.
Now we know that the weak interaction does not conserve parity. Hence
for +pz we have h = +1. Let us associate this with the left-handed neutrino.
As parity is not conserved in the weak interaction, h = +1 and h = −1
states should be independent states and thus the other neutrino – the right-
handed neutrino – should be an independent representation. There is thus a
subtle but fundamental difference between what is a left-handed neutrino and
what is a right-handed neutrino. Since only one of these manifests itself in the
laboratory, what is this other helicity partner of it? We studied this problem
above in terms of the independence of right-handed neutrino as a completely
different entity.

10.10 Solutions of Problems


Solution 10.1:
Yl
Yφ = 1 + δ where δ  1. Yl = −Y − δ and Yq = − Nc
= (Yφ + δ) /Nc . Thus
408 Group Theory in Particle, Nuclear, and Hadron Physics

 
u 1 1 (δ/Y φ)
qL = : Q (uL ) = + +
d L 2 2Nc 2Nc
1 1 (δ/Y φ)
Q (dL ) = − + +
2 2Nc 2Nc
1 1 δ/Y φ
uR : Q (uR ) = + +
2 2Nc 2Nc
1 1 δ/Y φ
dR : Q (dR ) = − + +
2 2Nc 2Nc
 
νe 1 1 δ
eL = : Q (νL ) = + (−Yφ − δ) = −
e L 2 2Yφ 2Yφ
1 1 δ
Q (eL ) = − + (−Yφ − δ) = −1 −
2 2Yφ 2Yφ
δ
eR : Q (eR ) = −1 −
2Yφ
 
1 3 3 (δ/Y φ)
Q(p) =
1+ +
2 Nc 2 Nc
 
1 3 3 (δ/Y φ)
Q(n) = −1 + +
2 Nc 2 Nc
————————————————–
Solution 10.2:

   
µ Yq 4 Yu
L (U (1), Quark) = q¯L iγ −ig1 Bµ qL + ūR iγ ig1 Bµ uR
2Yφ 2Yφ
 
¯ µ Yd
+dR iγ ig1 Bµ dR
2Yd

1 
Yq ūL γ µ uL + d¯L γ µ dR + Yu ūR γ µ uR + Yd d¯R γ µ dR Bµ
 
= −g1
2Yφ
g2  µ 0
√ µ +
√ µ − µ 0

Lf (SU (2), quark) = ūL γ uL Wµ − 2ūL γ dL Wµ − 2d¯L γ uL Wµ − d¯L γ dL Wµ
2

Lem = Q(u)Aµ (uL rµ uL + uR rµ uR ) + Q(d)Aµ dL rµ dL + dR rµ dR




Y Y
g2 Aµ + g1 Yφq Zµ −g1 Yφq Aµ + g2 Zµ
0
Bµ = p ; Wµ = p
g11 + g22 g12 + g22
Glashow-Salam-Weinberg Model 409

   
1 Yq g2 1 Yq g2
L = −uL rµ uL − g1 Bµ − Wµ0 − d¯L rµ dL − Bµ + Wµ0
2 2Yφ 2 2 Yφ 2
   
1 Yu 1 Yd
−uR rµ uR − g1 Bµ − dR rµ dR − g1 Bµ
2 Yφ 2 Yφ
Collecting u-terms yields
    
µ Yq Yq µ 1 Yu
= Aµ ((ūL γ uL )g1 g2 + + uR r uR + g1 g2
2Yφ 2Yφ 2 Yφ
2
L= Aµ (uL rµ uL + uR rµ uR )
3
And similarly for the d-term and Zµ as well.
————————————————–
Solution 10.3:

m20 c4 m20 c4 c2
= = −1
p2 c2 (m0 ϑv)2 c2 v2
r
c2 c2 c
vp = c 1 + 2 − 1 = =
v v β
And thus vp > c, while,
c
vg = q =v
2
1 + vc 2 − 1
————————————————–
Solution F.1:

eiKx ∆φ eiKy ∆ψ e−iKx ∆φ e−iKy ∆ψ


= 1 − [Kx , Ky ]∆φ ∆ψ + Kx2 (∆φ)2 Ky2 (∆ψ)2 + . . .
Now as [Kx , Ky ] = −iJz , we obtain an overall rotation around the z-axis.
————————————————–
Solution F.2:

[J~2 , K1 ] = [J12 , K1 ] + [J22 , K1 ] + [J32 , K2 ]


[J12 , K1 ] = 0
[J22 , K1 ] = J2 [J2 , K1 ] + [J2 , K1 ]J2
= −iJ2 K2 − iK3 J2
and [J32 , K1 ] = iJ3 K2 + iK2 J3
Hence in general [J~2 , Ki ] 6= 0
————————————————–
Chapter 11
Symmetry in Nuclei

11.1 Isospsin Symmetry in Nuclei . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 411


11.2 SU (4) Symmetry and Saturation in Nuclei . . . . . . . . . . . . . . . . . . . . . . 418
11.3 Distinguishable Protons and Neutrons in Nuclei . . . . . . . . . . . . . . . . 420
11.4 Gamow-Teller Strengths in Nuclei . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 423
11.5 SU (2)A Nusospin Symmetry in Nuclei . . . . . . . . . . . . . . . . . . . . . . . . . . 426
11.6 Quantum Groups in Nuclei . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 431

11.1 Isospsin Symmetry in Nuclei


Proton and neutron are spin 12 fermions, with baryon number B=1. How-
ever, they possess different charges. Their masses are mn = 939.6 M eV and
mp = 938.3 M eV . As both of these constitute the nucleus, we hypothesize
that they are “identical” particles or the “same” particle (commonly termed
the ‘nucleon’) occurring in two forms as proton and neutron. The idea is sim-
ilar to treating spin-up and spin-down as two forms of the same spin - 12 . This
duality of spin is extended to the nucleon with a physically different SU (2)I
group – denominated isospin to specify it. But the problem is that these
masses are slightly different. We assume that this difference in mass:
mass − dif f erence mn − mp
= 1 ∼ 10−3 (11.1)
mean − mass 2 (mn + mp )

is small and of an entirely electromagnetic origin. If so, the implication is that


if the electromagnetic force were to be turned off, there would be no difference
in their masses and then these would be simply two manifestations of the same
“identical” particle of the SU (2)I group – which would then likewise become
the proper definitional group of the nucleus.
With the knowledge that this effect on mass due to electromagnetism is
small, we take SU (2)I as a good approximate symmetry of the nucleus, with
(p,n) constituting the fundamental representation of SU (2)I . This hypothesis
is confirmed by the property of charge symmetry and charge independence of
the nuclear forces.
One may emphasise that the mathematical group SU (2) exists as an ab-
stract entity. It turns out nature makes use of this as a “word” to map the

411
412 Group Theory in Particle, Nuclear, and Hadron Physics

overall mathematical reality of the spin degrees of freedom of elementary en-


tities – let us denote it the group SU (2)S , where the subscript S signifies the
spin degree of freedom. It turns out that the same “word” finds application
in mapping another reality – the approximate isospin symmetry in nuclei. In
the socio-cultural framework of what constitutes a language, a word can only
be comprehended through its interactive relationship with other words (see
Appendix A).

Charge Symmetry
This property involves a pair of “mirror nuclei”. Mirror nuclei are a pair
of nuclei which are obtained from each other by the exchange of the number
of neutrons and protons with each other. This n ⇔ p connects 3 H and 3 He
nuclei as the simplest pair of mirror nuclei.
In 3 H there are 3 pairs of 2 body interactions: p-n, p-n and n-n while
in 3 He there are p-n, p-n, p-p. So the only difference is that one has n-n
interaction while the other has p-p. While n-n has only the strong part of
interaction, p-p has both the strong part and the electromagnetic part too.
Due to isospin symmetry being a source of the strong interaction, we assume
that (n-n) and (p-p) strong force should be identical. In that case the difference
in the binding energy of 3 H(8.492) MeV and 3 He(7.728) MeV should come
entirely from the Coulomb interaction of the first nucleus. We calculate with
1
R ∼ 1.45 × 10−13 A 3 fm,
e2
= 12 Z(Z − 1). 65 . 4πR
= 0.826 M eV
Now we assume that, given the fact that the formula for R is being utilised
for too light a nucleus here, this is still a good result to confirm the assumption
of (n-n) and (p-p) forces being equal to one other.
Energy levels of pairs of nuclei, like 11 11
6 C5 and 5 B6 are seen to match pretty
well starting with the ground state energy at zero. This equality of (n-n) and
(p-p) forces is termed the charge symmetry of the nuclear force.

Charge Independence
Above we saw that (n-n) and (p-p) forces in a nucleus are identical. Now on
account of the isospin symmetry when (p-n) are treated as identical particles,
we now have three pairs of identical particles: (n-n), (p-p) and (n-p). So if (n-
p) are in the same relative state of motion and are in the same singlet state, we
expect that – due to the isospin symmetry – this interaction should be similar
to the other two. Thus for a pair of nucleons N-N, the states are |T, To >:
|1, 1 >, |1, −1 >, |1, 0 >. Hence the equality of the nuclear interactions of
these states means that these will be degenerate in energy with respect to one
another.
As electric charge is related to the third component T0 of the isospin
operator T~ , the charge independence of the nuclear force implies the following
commutation:
Symmetry in Nuclei 413

[H, T0 ] = 0 (11.2)
where H is the nuclear Hamiltonian.
Hence the eigenstates of these H as |ψ > are simultaneously eigenstates of
T0 as well,
1
T0 |ψ >= (Z − N )|ψ > (11.3)
2
Thus the nuclear force is invariant under a rotation in isospin space. As H
also commutes with the isospin Casimir operator,
h i
H, T~2 = 0 (11.4)

whence

T~2 |ψ >= T (T + 1)|ψ > (11.5)


Thus as long as T remains unchanged, p and n can be exchanged to obtain
good isospin states which can be connected to each other as isobaric analogue
states (IAS) due to invariance of rotation in isospin space. Such IAS have been
well studied, confirming the property of charge independence of nuclear force.
This is understood as follows: Take two isobars, i.e., nuclei with the same
(Z−N )
mass number, for example, 21 21
11 N a and 12 M g. TZ = 2 is 12 and 32 , respec-
tively, for these nuclei. We assume these represent two states of isospin T= 32 .
As per the conception of charge independence, these states should exhibit the
same parity, spin and energy separations. These corresponding states are the
IAS states and possess similar wave function structures. The IAS are studied
by means of (p,n) reactions; the incident proton knocks out the less bound
target neutron in a (Z-1, N+1) nucleus to get captured to create the IAS of
the nucleus (N, Z).
This charge independence appears to hold pretty well in the nucleus. How-
ever there do exist particular instances where this symmetry is broken, and
this can be studied as a perturbation effect.
Let us now explore a curious implication of charge independence. Assume
the p-n force is caused by the exchange of a charged pion, π ± , the range of
the force R(pn) ∼ (m1 + ) and mπ+ = mπ− . If there were only π ± then a
π
simultaneous exchange of a π + π − pair would give the force R(nn) = R(pp) ∼
1 1 + −
(2mπ+ ) = 2 Rpn . This is a stronger force as π π will glue more strongly than
before. But charge independence demands that all these forces be of equal
strength. Hence the need for a π 0 to mediate this force R(nn) = R(pp) =
R(pn) ∼ (m1 0 ) = (m1 + ) . Consequently the three pions π + π − π 0 mediate the
π π
strong force with equal strength. Observe that charge symmetry in itself is
not powerful enough to obtain the π 0 exchange term.
The strong interaction in the nucleus conserves isospin. Hence:
414 Group Theory in Particle, Nuclear, and Hadron Physics

d + d →4 He + π 0 + π 0 and d + d →4 He + π + + π −
These conserved isospin reactions do indeed occur in the laboratory. Mean-
while, isospin non-conserving reactions like say

d + d →4 He + π 0
are not observed. Now consider reactions,

p + d →3 He + π 0 and p + d →3 He + π +
Here the total isospin of the LHS is |p, d >= |T1 T3 >= | 12 , + 12 >. On the
RHS we have to combine |3 He > | 12 , + 21 > and |π 0 > |1, 0 > and |3 H >
| 12 , − 12 > and |π + > |1, 1 > to yield the respective states on the LHS. Using
CG coefficients:

 1 1
  1 1

1 1 1 1 1 1 1 1
| , + >= 21 2 | , + > |1, 0 > + 2 2 | , − > |1, 1 >
2 2 +2 0 + 12 2 2 − 12 1 1
2 2 2

Thus the cross-sections are related to the probability of the above states
as follows:
1 1 2
 
2 1 2
σ(pd →3 He π 0 ) + 12 0 + 12 1
= = (11.6)
σ(pd →3 H π + ) 1 2 2
 1 
2 1 2
− 12 1 1
2

This agrees with the observations of experiments.

Generalized Pauli Exclusion Principle


Now the nucleus is made up of two types of fermions – protons and neu-
trons. So for the (p-p) and the (n-n) pairs, the Pauli Exclusion Principle should
apply separately. But what about (p-n) pairs? Under the SU (2)I symmetry,
the proton and neutron are identical particles with I = 12 , S = 12 . This was
a new idea introduced in nuclear physics in the early 1930’s. But what does
it mean in terms of the Pauli Exclusion Principle? We know that identical
fermions should possess a wave function completely antisymmetric under the
exchange of any pair of particle coordinates. Hence it may be presumed that
two nucleons (proton and/or neutron) should be antisymmetric under an ex-
change of any two of them. The degrees of freedom available are: the orbital
symmetry (i.e. location of the particle in SO(3) space); the spin degree of free-
dom in SU (2)S , and; the isospin symmetry of the two nucleons in the SU (2)I
group. Now the orbital states l = 0, 2, 4 are symmetric while l = 1, 3, 5 are
anti-symmetric; the spin state S=1 is a symmetric state and the spin state
S=0 an antisymmetric state; the isospin state I=1 is symmetric and the isospin
Symmetry in Nuclei 415

state I=0 is an antisymmetric state for the exchange of the two nucleons. The
total wave function of say a NN pair is:

Ψ = ψspace χspin φisospin (11.7)


3
Now the deuteron is mainly a S state, i.e., L=0, S=1. Moreover, as the
deuteron is deformed there is a small D-state admixture 3 D i.e. L=2, S=1
which in orbital space is still symmetric. Hence, in order for it to be fully
antisymmetric, the deuteron must be in an isospin I=0 state.
Deuteron has the ground state ( L=0, S=1, I=0 ) as the only bound state
with no excited states. Now the (n-p) pair can also exist in the isospin I=1
state as the I3 = 0 component with I3 = 1 for the (p-p) pair and I3 = −1 for
the (n-n) pair. Hence the I=1 state is not bound in a nucleus. Thus deuteron
binding demonstrates that the N-N interaction is spin dependent.
Now we know that

~σ1 · ~σ2 χm m
σ1 · ~σ2 χS=0 = −3χS=0
s=1 = χs=1 ; ~ (11.8)
And hence,

1 1
(1 + ~σ1 · ~σ2 )χm m
s=1 = χs=1 ; (1 + ~σ1 · ~σ2 )χs=0 = −χs=0 (11.9)
2 2
This is exactly the property of the spin-exchange operator,

σ 1
P12 = (1 + ~σ1 · ~σ2 ) (11.10)
2
which interchanges the spin variables of the two nucleons. Thus,
σ m
P12 χs=1 = χm σ
s=1 and P12 χs=0 = −χs=0 (11.11)
Or one has,
σ
P12 χs = (−1)S+1 χs (11.12)
This as triplet is symmetric and singlet is antisymmetric. For isospin group
similar exchange operator as in Equation 11.10

τ 1
P12 = (1 + ~τ1 · ~τ2 ) (11.13)
2
With similar exchange properties,
τ
P12 φT = (−1)T +1 φT (11.14)
the space exchange yields

pr12 ψl = (−1)l ψl (11.15)


416 Group Theory in Particle, Nuclear, and Hadron Physics

For historical reasons the spin, isospin and space exchange operators are
denominated Bartlett, Heisenberg and Majorana operators, respectively.
So as per the Generalized Pauli Exclusion Principle, in the exchange be-
tween any pair of nucleons in a nucleus, we demand that the total wave func-
tion be antisymmetric under the exchange of spin, isospin and space coordi-
nates. Consequently,
r τ σ
P12 p12 p12 Ψ = −Ψ (11.16)
And we can express this as,
r τ σ
P12 P12 p12 = (−1)L+S+T = −1 (11.17)

One-Pion-Exchange Potential (OPEP) :


Quantum mechanically it is known that l=0 is the lowest bound state
or ground state of a bound system. Thus one would expect that a bound
state of (n-p) should be an l = 0 state. However the quadrupole moment of
the deuteron has been measured to be Q = 0.286ef m2 . A spherical charge
distribution for l = 0 does not have a non-zero quadrupole moment. Only an
operator ∼ r2 can have a non-zero quadrupole moment, and this is zero for
spin − 21 objects, but nonzero for spin-1 object like deuteron. Meanwhile, the
L=0 state in J = L + S = 0 + 1 = 1 does not contribute to the quadrupole
moment. Only the J = ~2 + ~1 → ~1 + . . . state contributes from the L = 2 term
in deuteron ground state (see Chapter 9 for details).
Now as for the spherically symmetric potential with l = 0 in the ground
state, we require that the NN potential should have an appropriate non-
spherical, tensor interaction term, in order to induce the L = 2 ∼ r2 term
in the deuteron wave function. Thus we need to build up an operator of rank-
2, as the quadrupole moment operator is of rank 2. Our degrees of freedom for
the deuteron are spatial and spin. Such an operator exists and is denominated
the tensor operator. It is given by,
3
S12 = (σ~1 · ~r)(σ~2 · ~r) − σ~1 · σ~2 (11.18)
r2
where ~r = r~1 − r~2 and subscripts 1 and 2 label the nucleon number. This
operator has been studied in detail in Chapter 9. Here we wish to see where
such an operator arises from.
The charge of a nucleon is defined as,
1
Q = t0 + (11.19)
2
where t0 = + 12 for the proton and − 12 for the neutron. The charge of the pions
is Q = T0 where T = +1 f or π + , −1 f or π − and 0f or π 0 . The charge of a
nucleus is
Symmetry in Nuclei 417

A
Q = T3 + (11.20)
2
where T3 = (Z−N 2
)
and A = Z + N .
We can thus visualize quantum mechanical fluctuation of states p → n +
π + , n → p + π − , n → n + π 0 , p → p + π 0 . These also conserve the electric
charge. Thus ~π acts as a mediator of interaction between nucleons. This is the
well-known Yukawa meson exchange model.
Note that in the lowest order of approxmination we may assume that the
interaction between any two nucleons is of the form of an interaction between
two dipoles,
  
~1 σ~2 , ∇
V (r) ∼ σ~1 , ∇ ~2 f (r) (11.21)

Now from the Yukawa model of one-pion exchange we assume the radial
dependence to be of the form,
r
e− r0
f (r) = (11.22)
r
where
~c
r0 = (11.23)
mπ c2
Here (r = |r~1 − r~2 |) yields the range of the one-pion-exchange interaction.
We may denote the pion-nucleon coupling constant as g. We therefore write
the potential as
 − ro r

~2 e
 
V (r) = −g 2 r02 σ~1 · ∇~1 σ~2 · ∇ (11.24)
r
For any function φ(r) one can prove that
 2 
~ σ · ∇φ)
~ ∂ φ 1 ∂φ 1 ∂φ
∇(~ = r̂(~σ .r̂) − + ~σ (11.25)
∂r2 r ∂r r ∂r
which yields

"  −r #
g2 3r02

3r0 e ro 2
V (r) = 1+ + 2 S12 + σ~1 · σ~2 − 4πr0 δ(r)σ~1 · σ~2
3 r r r
(11.26)
where S12 is the tensor term above.

————————————————–
Unsolved Problem 11.1: Derive Equation 11.26 in detail.
————————————————–
418 Group Theory in Particle, Nuclear, and Hadron Physics

Now clearly this tensor term in OPEP induces an L = 2 (D-state) ad-


mixture in the deuteron ground state, which serves to explain its non-zero
quadrupole moment. It is interesting to observe that the lightest nucleus itself
is already deformed in the ground state. We already saw that the one-gluon
state exchange potential between quarks likewise possesses a tensor term which
would induce a D-state admixture in the ground state of a nucleon to produce
a deformation in the very ground state itself.

11.2 SU (4) Symmetry and Saturation in Nuclei


Looking at the systematics of the binding energy per nucleon in a nucleus,
one finds that for A > 30 it is fairly constant at ∼ 8 − 8.5 M eV . This means
that each nucleon on an average must be interacting with only a few nucleons
in its immediate vicinity.
For light nuclei up to and including 32 16 S, one notices that there occur sharp
peaks for nuclei A = 4n at n = 1.2, 3..... All these are Z = N even-nuclei:
4 8 12 16 20 24 28 32
2 He, 4 Be, 6 C, 8 O, 10 N e, 12 M g, 14 Si, 16 S. For nuclei next to these,
the binding energies per nucleon are much smaller than for the above nuclei.
Hence for 32 H, 31 He, 52 He, 53 Li these are 2.83M eV, 2.57M eV, 5.48M eV and
5.48M eV , respectively . For 4 He it is a substantially higher 7.07 MeV, as
is true for all the above (4-n) nuclei. This property – that for the nucleus
built up of α− nuclei the interaction is much stronger than in the neighboring
nuclei – is termed the saturation of the nuclear forces. The interaction energy
between the α-pairs in these nuclei is relatively weak; e.g., 8 Be is unstable to
decay into two α0 s. It turns out that on an average ∼ 90% of binding energy is
taken away by nucleons located in separate α0 s and about ∼ 10% by nucleons
in different clusters.
It is commonly believed that this saturation property in a nucleus is a
reflection of the fundamental symmetry SU(4) or, as it is popularly known,
the Wigner Supermultiplet symmetry [[98], page 163].
Now nucleons belong to the SU (2)I isospin symmetry group. Moreover,
each nucleon additionally possesses a spin degree of freedom specified by the
SU (2)S symmetry group. Together these provide four internal degrees of free-
dom. Thus if the nuclear interaction, in addition to the traditional indepen-
dence of charge, is also independent of spin, then one may enlarge the internal
group structure to a dynamical symmetry SU (4)IS ⊃ SU (2)I ⊗ SU (2)S and
assume that its fundamental representation is:
 ↑
p
 p↓ 
 ↑
n  (11.27)
n↓
Symmetry in Nuclei 419

Thus, 4 He in this representation is spin 0, isospin 0 and is identified with


the fundamental representation of SU (4)SI . Similarly all 4n, n = 2, 3 . . . nuclei
would have spin 0 and isospin 0 and hence be extra stable. These could be
treated as composites of only α’s. Thus the saturation property of the nucleus
is naturally explained here.
Remember that just as 4 He is treated as composed of four elementary
particles - two protons and two neutrons, here we are treating the nucleus
16 4
8 O as being constituted of four ‘elementary’ He’s.
However there may be problems in this interpretation: firstly, as it is only
the total angular momentum which has been observed experimentally, the
assignment of spin S=0 may be doubtful; and secondly, the drastic assumption
that the nucleon interactions are isospin and spin independent may not be
founded on reality.
However from N-N scattering itself, it is observed that the potential ex-
hibits spin and isospin dependence. Moreover, the deuteron as an n-p sys-
tem possesses isospin 0 and spin 1. The (n-p) system also has another state
with (T = 1, T2 = 0) and spin 0, and this one is not bound. Thus there is a
strong spin-dependence here in the N-N interaction. Thus as spin-dependence
is clearly apparent even in the 2-N system, it is not reasonable to expect the
4-N system to suddenly become spin independent. Hence the above group the-
oretical model prediction does not hold good physically. As experiments are
the ultimate arbiters in physics, perceived ‘beautiful’ mathematical structures
should not mislead into wrong science.
So what is the correct explanation for the nuclear saturation property?
Now it has also empirically emerged from NN scattering that there exists
strong repulsive core (i.e. hard core) for relative distances between nucleons
less than ∼ 0.5 f m. It is believed that this repulsive core exists in all states
and that it is important in explaining the saturation of nuclear forces [99].
Thus one starts by describing the nucleus as built of various proton-neutron
states, but truncating them or cutting them off at short distances due to the
hard core. This is an additional physical effect essentially put in by hand.
Hence the origin of the hard core clearly lies outside the purview of mere
nucleonic degrees of freedom. This, we shall show in the next chapter, is a
phenomenon arising from the quark effect in nuclei. Hence we shall demon-
strate that the saturation of nuclear forces is actually a manifestation of quark
degrees of freedom in nuclei.
One important point should be noted. Here, when we extend the SU (2)I
group (with (p,n) providing its fundamental representation) with SU (2)S to
a larger dynamic group SU (4)S+ we find that it fails to provide a reasonable
description of the simplest A=2 nucleonic system. However as we shall show
in Chapter 7, when the SU (3)F group with (u,d,s) quarks providing its funda-
mental representation is extended – through means of the incorporation of the
spin SU (2)S group to form a dynamic group SU (6)SF – then it works very
well to specify all the hadrons. This is rather surprising given that the smaller
group SU (2)I fails against a larger group SU (3)F when extended to a larger
420 Group Theory in Particle, Nuclear, and Hadron Physics

dynamical group. Which shortcoming does the group SU (2)I – as applied to


nuclei – suffer from, which is absent in SU (3)F as applied to hadrons?

11.3 Distinguishable Protons and Neutrons in Nuclei


It has already been shown how the SU (2)I group – which treats each pair
of neutrons and protons in nuclei as indistinguishable (all identical) fermions,
subject to the Generalized Pauli Exclusion Principle – gives a very successful
model of the nucleus. The only evidence of any uncomfortable fact for the
SU (2)I isospin group is the failure of the extended SU (4)SI ⊃ SU (2)I ⊗
SU (2)S group to describe the property of nuclear saturation in the nucleus.
We trace this deficiency of the SU (4)SF to the following facts.
It may come as a surprise, specially to the younger generation (whose
knowledge of nuclear physics may be limited to the nevertheless successful
SU (2)I isospin model), that there exists an equally successful model of the
nucleus wherein protons and neutrons are treated as distinguishable fermions.
This means, as to the fermionic aspect, protons and neutrons are as different
from each other as electrons are from protons. So in this picture the Pauli
Exclusion Principle is separately valid for a (p-p) pair and for a (n-n) pair, but
not for a (n-p) pair. And this model for distinguishable protons and neutrons
is as successful as the SU (2)I model of the nucleus. It is amazing that two
different ways of applying the Pauli Exclusion Principle to the same set of
fermions produce equally successful descriptions of them. Should we not call
it a fundamental duality?
In fact, right up to the ∼ 1960s most of nuclear physics was undertaken
by treating protons and neutrons as distinguishable fermions. This also holds
for a few more recent works in nuclear physics. Here we refer to Blatt and
Weisskopf [100].
However the fact is that a nucleus can be described well in an SU (2)I
model (where (p-n) are distinguishable) and in another independent picture
where the pair (p-n) is treated as made up of distinguishable fermions. Lawson
has shown, in a complete section entitled “Isospin and non-isospin methods of
calculation”, that these two independent methods yield essentially identical
results [101].
First he calculated the spectrum of 35 17 Cl by using the isospin formalism
and after this he obtained similar results utilising the neutrons-protons for-
malism (where they are treated as distinguishable particles). Also similarity
and success of these two independent methods was repeated in the study of
the spectra of 39
18 Ar. Note that these amazing results are highly non-trivial in
nature.
The relationship between the two formalisms here is discussed at many
places [100], [99]. Below we shall follow the notation of the former.
Symmetry in Nuclei 421

Let us first treat the protons and neutrons as different (i.e. distinguishable)
fermionic particles in a nucleus (~r is the space coordinate and χ is the spin
coordinate). Now,

φ = φ(r~1 χ1 , r~2 χ2 , . . . r~N χN ; ~rN +1 χN +1 , . . . r~A χA ) (11.28)


where up to N are coordinates of neutrons and the rest of protons. The Pauli
Exclusion Principle requires that the wave function changes sign if any two
protons are exchanged or if, independently, any two neutrons are interchanged.
There is no special requirement as to the exchange of protons and neutrons
in φ.
Next let us introduce a new isospin coordinate η = +1 and − 1,

ψ = ψ(~rχ1 η1 , r~2 χ2 η2 , . . . r~A χA ηA ) (11.29)


Note that the above ψ vanishes unless η1 = η2 = · · · = ηN = +1 and
ηN +1 = · · · = ηA = −1 , and lets the wave function change sign when all
the coordinates - position, spin and isospin are interchanged. This selection
changes sign under interchange of the first N-coordinates or within the other
set of (A-N) coordinates. But it does not satisfy this condition for an inter-
change of say r1 χ1 η1 with rA χA ηA coordinates.
Now the ordinary wave function of a two-neutron system Equation 11.28
is also antisymmetric under the interchange of these two neutrons,

φ = φ(r~1 χ1 ; r~2 χ2 ) = −φ(r~2 χ2 ; r~1 χ1 ) (11.30)


Next let us introduce the isospin wave function for a neutron,

ν(η) = 1 if η = +1
ν(η) = 0 if η = −1 (11.31)
Thus this wave function in two formalisms, Equation 11.29 and the above,
is

ψ(r~1 χ1 η1 , r~2 χ2 η2 ) = φ(r~1 χ1 ; r~2 χ2 )ν(η1 )ν(η)2 ) (11.32)


This is clearly antisymmetric due to the antisymmetry of φ.
For the proton we similarly define,

π(η) = 0 if η = +1 (11.33)
π(η) = 1 if η = −1 (11.34)
Now φ does not change sign under 1 ↔ 2 exchange. So include the isospin
and we have two possible states,

ψI = φ(r~1 χ1 , r~2 χ2 )ν(η1 )π(η2 ) (11.35)


and
422 Group Theory in Particle, Nuclear, and Hadron Physics

ψII = φ(r~2 χ2 , r~1 χ1 )π(η2 )ν(η1 ) (11.36)


Now quantum mechanically we cannot determine which particle is at po-
sition no. 1 and which at position no. 2. Hence,
1
ψ(~r1 χ1 η1 , r~2 χ2 η2 ) = √ (ψI − ψII ) (11.37)
2
1
= √ {φ(~r1 χ1 , ~r2 χ2 )ν(η1 )π(η2 ) − φ(~r2 χ2 , ~r1 χ1 )π(η2 )ν(η1 )} (11.38)
2
And as required in the isospin formalism, it is antisymmetric under 1 ↔ 2
exchange.
Hence this demonstrates that it is merely a formal requirement to move
from one formalism to another. So taking the Pauli Exclusion principle for
the proton and neutron separately in a conventional manner or by requiring
antisymmetry under the exchange of two nucleons in isospin formalism, i.e.
no matter whether we had (p − p) or (p − p) or (n − p) pairs, we are able to
build an antisymmetric wave function from the conventional wave function φ.
There is a subtle difference though. As was shown above, every conven-
tional wave function like φ can be generalized to be written in the proper
isospin formalism. But the converse does not always hold. Take the case of a
simple isospin wave function of the single nucleon,
1
ψ(~r, χ, η) = φ (~r, χ) √ {ν(η) + π(η)} (11.39)
2
This function corresponds to a nucleon in the ordinary states φ(~r, χ). How-
ever, this nucleon has equal probability of being a proton or a neutron at any
particular time. This state does not correspond to any physically known states
of a proton or a neutron. Thus the isospin formalism provides us with spuri-
ous states which do not correspond to any physical reality whatsoever. Hence
the need to use proper states in the isospin formalism, which means that we
ensure that states with a definite number of protons and neutrons only are
constructed.
So the two frameworks as independent methodologies do provide a satis-
factory description of the nucleus. Hence there exists a very fundamental du-
ality in the description of nuclear physics reality. This Manichaean character
should not be too shocking, as it bears a sharp resemblance to the analogues
wave-particle duality of entities in quantum mechanics.
Thus we are forced to treat the neutron-proton system in different ways
and still the final physical quantities are insensitive to this difference. Hence, in
addition to the isospin formalism, there should be another group theoretical
structure wherein the proton and the neutron are taken as distinguishable
particles. In that case there should not be any identifiable isospin quantum
number labelling neutrons and protons. This is just like in case of the electron
and the proton, in that both being fermions, they are actually distinguishable
Symmetry in Nuclei 423

because of their distinct quantum numbers. So in which model are protons


and neutrons non-isospin partners and what does group theory have to say
about it?
As we shall discuss in Chapter 7, the spin − 12 members of a baryon octet
are described as simultaneously made up of a 3-dimensional representation
of quarks in the SU (3)F quark model, (3 ⊗ 3 ⊗ 3 = 1 ⊕ 8 + 8 + 10) and also
independently as an adjoint representation of the same group SU (3)F . Now
as per the formal case of the SU (3)F quark model we take the proton and the
neutron as being treated as identical particles in the SU (2)I isospin subgroup,
and hence for a (p-n) antisymmetric wave function if needed. This leads to the
formulation of the Generalized Pauli Exclusion principle in nuclei. However
as per the second case of belonging to the adjoint representation of SU (3),
we find that the protons and neutrons do not have isospin quantum numbers
to define them. Hence in this formulation, they are not identical fermions
and should be treated as distinguishable fermions. Thus the duality of the
description of the (p,n) fermions in the SU (3)F octet model and that in the
adjoint representation of the octet model is reflected as a one-to-one dual
description of the nucleus as specified above.
Another interesting aspect which demands that we go beyond nuclear
isospin is the presence of “effective charges” in nuclei. Proton has electric
charge of one unit and neutron is neutral. But to understand the electromag-
netic experimental quantities in its wholeness in the nucleus, phenomenolog-
ically neutron has half a unit of electric charge and proton has an additional
half a unit of electric charge [102]. Is this an indication of neutron and proton
being distinguishable particles?

11.4 Gamow-Teller Strengths in Nuclei


Beta decay of a free or a bound neutron may occur:

n → p + e− + ν̄e (11.40)

A − −
Z XN →A
Z+1 X N −1 + e + ν̄e (β decay) (11.41)
Because a neutron is heavier than a proton (939.57 MeV to 938.27 MeV) a
proton cannot undergo a corresponding beta decay. However a bound proton
is provided with extra energy in a nucleus and it’s beta decay may occur:

(p)bound → n + e+ + νe (11.42)

A
Z XN →A + +
Z−1 AN +1 + e + νe (β decay) (11.43)
424 Group Theory in Particle, Nuclear, and Hadron Physics

The free neutron decay (Equation 11.40) half-life is ∼ 10.37min. If (e− +


ν̄− ) pair is in a total spin zero state then the proton is in the same spin state
as the neutron. Thus the operator which transforms the neutron spin to the
proton spin is a unit operator 1. Such a decay – where only the flavour changes
(n to p) with no change in the spin – is termed the Fermi decay. However the
(e− ν̄e ) pair in spin states one (triplet-spin) is denominated the Gamow-Teller
(GT) decay. Here the spin operator ~σ (in addition to isospin ~τ ) is required to
excite the spin degree of freedom. Now compare ~σ to 1 in spin space, and we
expect, that as < ~σ 2 >∼ 3, the GT decay should be three times as significant
2
as the Fermi. However it experimentally turns out that this number is 3gA :1
where gA = 1.25 represents a renormalization of the weak interaction in the
GT decay. This indicates the presence of an internal structure of the nucleon
and this we shall see in the quark model.
As τ+ |n >= p etc. the operator for the β − decay in a nucleus is as follows.
Fermi decay:
A
X
T± = τ± (i) (11.44)
a=1

Gamow Teller decay:


A
X
Y ± = gA ~σ (i)τ± (i) (11.45)
a=1

Form Equation 11.44 as τ+ |n >= |p >, τ+ |p >= 0, τ− |p >= |n > in


Fermi transitions the strength (Equation 11.44) is:

SβF− − SβF+ = N − Z (11.46)


in a charge exchange reaction of (p,n) type,

A(Z, N ) + p → B(Z + 1, N − 1) + n (11.47)


Here the nuclear structure component is the same as in β − decay. The
prime advantage of the (p,n) reaction is that it describes the above structure
of the β ± decay operator without worrying about the Q-value relevant for free
neutrons and protons. However, the (n,p) reaction is more difficult to study
as it is not easy to obtain neutron beams in the laboratory. Yet it describes
the β + counterpart of Equation 11.47 above.
Let us calculate the corresponding sum-rule for the GT strengths as:

X A
X 2
SβGT GT
− − Sβ + = |< n| gA ~σ (i)τ− (i) |0 >|
n i=1
A
X A
X 2
− |< n| gA ~σ (i)τ+ (i) |0 >| (11.48)
i=1 i=1
Symmetry in Nuclei 425

Writing the spin operator in spherical representation and noting that the
above strength may be written as < 0 |[β+ , β− ]| 0 >

XX XX
SβGT GT 2
− − Sβ + = gA < 0|[ (−1)µ σµ (i)τ− (i), σµ0 (j)τ+ (j)]|0 >
µ i µ0 j

XX X X
2
= gA < 0| [ (−1)µ σ−µ (i)τ− (i), σµ0 (i)τ+ (i)]
µ µ0 i i
X X
+[ (−1)µ σ−µ (i)τ− (i), σµ0 (j)τ+ (j)] (11.49)
i6=j i6=j

The second term is zero. So,

XXX
SβGT GT 2
− − Sβ + = gA < 0| (−1)µ σ−µ (i)σµ0 (i)[τ− (i), τ+ (i)]|0 >
µ µ0 i

We know τ− τ+ − τ+ τ− = −τz and so,


A
X
SβGT GT 2
− − Sβ + = 3gA < 0|
2
τz (i)|0 >= 3gA T r(τz )
i=1

whence,

SβGT GT 2
− − Sβ + = 3gA (N − Z) (11.50)
We can easily obtain the individual strengths though. For example,
A
X
SβGT 2
+ = gA < 0| (−1)µ σµ (i)τ+ (i)σµ (i)τ− (i)|0 >
i=1

A
X X
2
= gA < 0| ~σ 2 (i)τ+ (i)τ− (i)|0 >= gA
2
[3 < 0| ~σ 2 (i)|0 >]
i=1

SβGT 2
+ = 3gA Z (11.51)
and similarly

SβGT 2
− = 3gA N (11.52)
These yield the differences as we obtained in Equation 11.50. So why bother
about the sum rule of the differences if we have predictions for individual
strengths themselves? The reason is that the individual strengths in Equations
11.51 and 11.52 represent the total strengths of all the protons and all the
neutrons excited by (n,p) and (p,n) reactions, respectively. These may involve
very high energy components. However only excitations near the Fermi surface
426 Group Theory in Particle, Nuclear, and Hadron Physics

may be participating more easily. But we have difficulties in understanding


these in a nucleus. However the difference of the two strengths may not be
sensitive to this problem and one may go to as high an energy as possible and
thus this prediction would be physically testable.
In fact departure from the sum rule in Equation 11.50 may indicate the
presence of nucleonic excitations like ∆ of the nucleon itself. If these ∆ – exci-
tation may be identified directly, then these will provide a new understanding
of the nucleus, and thereby the significance of the (p,n) and the (n,p) reac-
tions in nuclei. In fact this difference in GT strength may very well be studied
within the quark effects in nuclei, as we shall see in Chapter 12.
The GT strengths have been extracted through excitations of the so-called
giant resonances, which are collective excitations of the nucleus in reactions
like 90 Zr(p, n)90 N b. A large number of such studies have been performed, in
particular by (p,n) reactions. It turns out, as a generic property in all nuclei,
that the GT strength extracted in these reactions is always quenched by 40-
60% of the above GT sum-rule. It appears as if this is a fundamental property
of the nucleus and hence requires consistent study of the same. Several studies
of the GT strength have been conducted, including those that use tensor terms
which can distinguish between GT strengths sitting at ∼70 MeV verses those
at lower energies ∼ 20 MeV [103].
However the possible excitations of ∆ in nucleons as observed through
GT strengths, are best studied by taking the quark structure of baryons in a
consistent manner in a nucleus.

11.5 SU (2)A Nusospin Symmetry in Nuclei


We have studied how SU (2)I , with (p,n) providing its fundamental repre-
sentation, has proven to be a useful symmetry in studying a nucleus. Moreover,
the α cluster of the nucleus was shown is not consistently explained by the
SU (2)I group model. We have to resort to the quark degrees which then ex-
plain the hard-core structure of the N-N interaction and hence α particle too.
Are there any other cluster structures evident in nuclei?
Recently it was demonstrated that the A=3 cluster structure is very promi-
nently evident in several experimental studies [104]. We shall show below that
this is actually an experimental confirmation of a clear theoretical prediction
arising due to the occurrence of a new SU (2)A – nusospin symmetry in nuclei.
Just as it is assumed that (p,n) provides the fundamental representation of
the SU (2)I isospin group, inthe same manner, it is hypothesised that the
helion 32 He and triton 31 H as an (h,t) pair provides the fundamental rep-
resentation of a new SU (2)A – nusospin group in nuclei (note that we use A
to distinguish it from ‘A’ which stands for the mass number of the nucleus)
[36]. If the existence of protons and neutrons justifies the SU (2)I isospin sym-
Symmetry in Nuclei 427
TABLE 11.1: Inter-triton cluster bond
energies of neutron-rich nuclei
Nucleus n m EB − 8.48n (Mev) C (Mev)
9
3 Li6 3 3 19.90 6.63
12
4 Be 8 4 6 34.73 5.79
15
5 B 10 5 9 45.79 5.09
18
6 C 12 6 12 64.78 5.40
21
7 N 14 7 15 79.43 5.29
24
8 O 16 8 18 160.64 5.59

metry in the nucleus, then the physical presence of h- and t- as evidenced


by the large number of empirical tests [105] then may be used as the basic
justification for the SU (2)A nusospin group. Let us now justify the existence
of the SU (2)A nusospin model and discuss as to which physical considerations
justify this symmetry structure.
i) Binding energy and rms radii of 32 He and 3 H are (7.7 MeV,1.88 fm) and
(8.48 MeV, 1.7 fm), respectively. That of 4 He is (28 MeV at 1.67 fm). As much
as α clusters are so strongly bound and so compact, this provides a natural
explanation as to why such clusters may exist in nuclei. Note, however that
the above A=3 nuclei are pretty well compact too, especially as each nucleon
itself has a size of ∼ 1 fm. In addition, as we shall see, Chapter 12 3 He,3 H
and 4 He all have a “hole” at the centre of their density distribution. This in
itself brings in rigidity in their structure too. To understand this property, let
us treat these h- and t- clusters as existing inside nuclei.
It is well known that light nuclei with N=Z structure A=4-n,
n=1,2,3,. . . may be treated as made up of n number of α clusters. Next we
assume that all neutron-rich nuclei of structure Z 3Z H2Z can be assumed to be
composed of N=Z number of triton clusters 31 H2 . As each triton has a binding
energy of 8.48 MeV, the binding energy of these nuclei may be written as:

E = 8.48n + Cm (11.53)
where these n-clusters of tritons form m-bonds where C is the inter-triton
bond energy. Just as 16 8 O8 may be treated as having six bonds if it is made up
of 4 α clusters (sitting on the edges of a tetrahedron), we take 124 Be8 as made
up of four tritons sitting on the edges of a tetrahedron and having six bonds.
We list these nuclei in Table 1. The number of bonds arising from standard
geometric structure are well studied in α- nuclei [106].
Note that the average inter-triton bond energy is approximately 5.4 MeV.
Clearly this justifies the symmetry structure of these neutron-rich nuclei as
being made up of tritons only. If α in A=4-n (n=1,2,3 . . . ) nuclei makes
physical sense then similarly so should 3z 3
z X2z = z1 H2 as well.
ii) Clearly as the above triton clusters in nuclei manifest themselves lucidly,
these should be more strongly bound and hence compact compared to the
428 Group Theory in Particle, Nuclear, and Hadron Physics

surrounding nuclei. This point is similar to the fact that α nuclei are more
bound then the neighbouring nuclei. Actually these are magic nuclei, more
strongly bound than their neighbours. Hence if one or more neutrons is added
to the new magic nuclei, they will be expected to be more loosely bound than
the core nuclei, and thus these may be expected to form a loose halo structure.
So the two-neutron halo structure of 11 11 9
3 Li as made up of 3 Li8 = 3 Li6 + 2n
9
where 3 Li forms a compact core and the extra two-neutron has a halo structure
as found in experiments. So 14 4 Be10 would be a two-neutron halo built on a
compact triton-based core 12 4 Be. This model predicts all the nuclei which may
be halo-like, and the same is confirmed like 19 C and 20 C being one-neutron
and two-neutron halo nuclei, respectively. Thus this triton cluster model is
successful in defining the halo structure of one neutron and two neutrons
[107].
iii) Magic nuclei are those which are more stable and more strongly bound
than the adjoining nuclei in the N and the Z numbers. Already it was pointed
out [108] that there is a tendency in neutron-rich heavy nuclei, for new msgic
numbers, besides the standard shell model ones, to pop up out of the blue.
SU (2)A nusospin, predicts, as an extension of the above arguments, that for
neutron-rich nuclei, the new magic numbers of nuclei would be (Z, N ) pair of
(4, 8), (6, 12), (8, 16), (10, 20), (11, 22) and (12, 24). Note that the magic num-
bers for stable nuclei are in general N ∼ Z (true for light nuclei) and these
magic numbers are N = Z = 2, 8, 20, 40. . . . This change in magic numbers as
we move away from N ∼ Z nuclei to N  Z nuclei like the 3Z Z X2Z nuclei,
is due to strong triton clustering effects and which is predicted from SU (2)A
nusospin symmetry [105].
iv) The removal of a neutron from a nucleus is defined as the difference
between the binding energy of the nucleus of the neighboring one with one
neutron less.

Sn (Z, N ) = EB (Z, N ) − EB (Z, N − 1) (11.54)


And similarly for the proton separation energy as,

Sp (Z, N ) = EB (Z, N ) − EB (Z − 1, N ) (11.55)


As is well-known, a plot of these with respect to N and Z in nuclei show
sharp peaks locally with respect to the neighbouring nuclei at the magic num-
bers in N and Z. In fact this is how the new magic numbers discussed above
were discovered.
Treating all 3Z
Z X2Z nuclei as being bound states of Z-number of tritons
3 12
H
1 2 , Be one gets 4 − t,27 F is 9 − t etc. Viewed this way the relevant triton
degrees of freedom are being treated as ”elementary”. Let us pick up and
knock out a single triton (as one elementary unit) in one shot from these
bound states of tritons. We know of several experiments which do this. Now
we define a triton separation energy as
Symmetry in Nuclei 429

St (Z, N = 2Z) = EB (Z, N ) − EB (Z − 1, N − 2) − EB (t) (11.56)

Here an extra triton binding energy term is needed [109].


Let us plot St as a function of the number of tritons existing in a nucleus in
Figure 11.1. Here the binding energies are the experimental binding energies.
Thus t=5 corresponds to the nucleus 15 24
5 B and t=8 to the nucleus 8 O.
One clearly sees the odd-even effect here. Thus when the triton number is
even, the triton separation energy is significantly higher than the correspond-
ing one of the next adjoining odd-triton nuclei. This feature is similar to the
odd-even effect observed in one-neutron and one-proton separation energies
plotted with respect to neutron and proton numbers, respectively. There this
odd-even effect is attributed to the identical particle n-n and p-p pairing being
favoured in n-even or p-even nuclei, respectively. Thus for the triton separa-
tion energy case too the odd-even effect should be attributed to a t-t pairing
effect. In the conventional case the n-n and p-p pairing is attributed to a shell-
effect in a bound potential. So here too we should associate the t-t pairing to
a shell effect of a binding potential. Hence we are being forced to assume that
for these neutron-rich nuclei there should exist binding potentials where the
tritons are bound as “elementary” entities. What a binding potential does for
a bound nucleus as made up of protons and neutrons, is clearly what this new
potential is doing for a bound system of elementary tritons.
Note also the prominent feature of the highest peak in St versus t at t=8.
This is for the t=8 nucleus 24 16
8 O16 , just as 8 O8 in the standard shell model is
made up of 8-protons and 8-neutrons. Hence as the number eight is a magic
number, this means that this number fully closes a particular shell of the
binding potential.
The next-most prominent feature in Figure 11.1 pertains to that of the
highest peak in the separation energy for Nt =8 i.e. for 24 O. This is an equally
sharp dip for Nt =9, i.e., for 27 F .
We know that such drops in one-neutron and one-proton separation ener-
gies when moving from one Z/N number to the next is also an unmistakable
signal of the magicity character of a particular Z/N number. In the context
of our discussion here, magicity means a much stronger binding for a particu-
lar number of tritons as compared to the adjoining number of tritons. Hence
clearly here Nt =8 is a magic number with respect to different bound states
of tritons. So evidently there exists a shell structure of the bound states of
tritons wherein there is a large extra stability for Nt =8, thereby indicating
magicity for this nucleus.
Let us treat this triton binding potential to be of a Harmonic Oscillator
(HO) kind. In an HO potential the magic numbers are 2,8,20,40, etc. For
our bound state of tritons Nt =2 is where the system starts and hence may
be justifiably treated as a magic number. Next Nt =8, as indicated above, is
indeed a magic number. Unfortunately the data does not exist to Nt =20. But
430 Group Theory in Particle, Nuclear, and Hadron Physics
40
S2t
St

31

St/2(McV) 20

10

0
0 2 4 6 8 10 12 14 16
triton number

FIGURE 11.1: One and two triton separation energies as a function of triton
number

clearly as per our model here we predict that 60 Ca, as a bound state of twenty
tritons, would be a magic nucleus.
For the case of magicity in one-neutron/proton separation energies, to
avoid the jumps due to the odd-even effect, one resorts to the smoother plot
of two neutron/proton separation energies. Herein magicity is indicated by
kinks in the plot at appropriate neutron/proton numbers. Thus we also plot
two triton separation energies S2t as a function of the number of tritons in Fig
1. The kink at Nt =8 is most prominent, thereby justifying the magic character
of Nt =8. This novel aspect of new magic numbers of new shell structures as
justified by the nusospin symmetry SU (2)A needs to be fully explored in
future.
Note that as to single nucleons, we do have a deuteron as a bound system
of an (n-p) pair. But the (n-n) is unbound. Also tri-neutrons or tetra-neutrons,
in spite of intense experimental searches, have not been observed in the lab-
oratory. So though the light bound systems of neutrons do not exist, we do
believe that astrophysically stars of a size of about 10 km, and composed of
neutrons only, do exist as so-called neutron stars. On the other hand here we
notice that bound states of light clusters of tritons do exist as per the nusospin
SU (2)A group. Thus it would be logical to expect that there may actually be
stars made up entirely of tritons only. This is an interesting hypothesis arising
from the new nusospin group.
Symmetry in Nuclei 431

11.6 Quantum Groups in Nuclei


We have seen that the Lie groups and the Lie algebras were, as mathemati-
cal structures, developed during the end of the 19th century. Their application
to physics commenced essentially in the 1920’s. Hence it was a well-developed
mathematical structure before it found applications in physics.
Quantum groups are remarkable mathematical structures which have
emerged as mathematical abstractions arising from physical problems them-
selves: quantum inverse scattering theory [110] and solvable statistical me-
chanical models [111]. Though the word is often used to describe this new
mathematics as the “quantum groups”, however more properly the quantum
group is actually a “quantum algebra”. It is a deformation of the universal
enveloping algebra of the underlying classical group.
One may write the continuous deformation parameter as q = e~ so that as
~ → 0 we obtain the “classical” undeformed algebra. It is in this manner that
one may justify the use of the word “quantized” here.
In the last two decades, the interest in quantum groups and their applica-
tion in physics has grown substantially [112]. Quantum groups and quantum
algebras become relevant in physics where the limits of the applicability of Lie
groups and algebras are stretched. They describe perturbations from some un-
derlying symmetry structures such as quantum corrections and anisotropies.
Remember that the Lie group/algebra specifies certain intrinsic symme-
tries of a system. Hence we expect that “quantum” groups may take account of
deviations from these symmetries of the physical systems from the “classical”
Lie groups.
Here we study the application of the quantum group SU (2)q , the quantum
group deformation of the classical Lie group SU(2).
One defines the SU(2) Lie algebra in terms of the generators J+ , J− and J0
which satisfy the commutation relations

[J0 , J± ] = ±J±

[J+ , J− ] = 2J0 (11.57)


where J± = J1 ± iJ2 . The Casimir operator is
1
J~2 = (J+ J− + J− J+ ) + J02
2
= J− J+ + J0 (J0 + 1)
= J+ J− + J0 (J0 − 1) (11.58)
The quantum algebra SUq (2) is defined by the generators J+ , J− and J0
which satisfy the commutation relations:

[J0 , J± ] = ±J±
432 Group Theory in Particle, Nuclear, and Hadron Physics

[J+ , J− ] = [2J0 ] (11.59)


where [2J0 ] is a deformed q-number, with J0† = J0 and †
J+ = J− , and the
q-number,

q 2J0 − q −2J0
[2J0 ] = (11.60)
q − q −1
Hence the q-number is defined as:

q x − q −x
[2J0 ] = (11.61)
q − q −1
with the parametrization

q = et ( with t real ); [x] = sinh(tx)/sinh(t) (11.62)


and for

q = eit ( with t real ); [x] = sin(tx)/sin(t) (11.63)


As q → 1(or as; t → 0) the q-numbers become the standard number,

Lim(q→1) [x] = x (11.64)


and so the deformed q-algebra goes over to the classical algebra,
q→1
SU (2)q −−−→ SU (2) (11.65)
Note:

[0] = 0, [1] = 1, [2] = q + q −1 , [3] = q 2 + 1 + q −1 (11.66)


[−x] = [x], [x]q = [x](q−1) (11.67)
[n]! = [1][2]..[n][n − 1] . . . [2][1], [0]! = 1, [−n] = ∞ at n = 1, 2, 3, . . . (11.68)
Let us now understand as to what does the deformation of the SU(2) Lie
algebra mean. For SU (2)q since J± = Jx ± Jy , J0 = Jz we can write the
commutation relations of the deformed SU (2)q algebra as
i
[Jx , Jy ] = [2Jz ], [Jy , Jz ] = iJx , [Jz , Jx ] = iJy (11.69)
2
This is the q-analogue of the SO(3) algebra as SO(3)q . The classical group
SO(3) is obtained in the limit q → 1.
In the classical group SO(3) all the three directions, x-,y-, and z- are equiv-
alent to each other. But we see that in SO(3) the z-direction is no longer
equivalent to the x- and the y-directions. This is also true for SU (2)q . There-
fore it is not unreasonable to expect that SU (2)q may be better than SU(2)
in the description of the fundamental properties of objects deformed in the
z-direction, such as the diatomic molecules and the deformed nuclei.
Symmetry in Nuclei 433

As in the usual Lie algebra mathematics the irreducible representations


D(j) of SUq (2) are determined by the highest weight states with j=0,1/2,1,...
The basis states |j, m > {−j ≤ m ≤ j} are connected to the highest weight
states as follows:
s
[j + m]!
| j, mi = (J− )j−m | j, ji (11.70)
[2j]![j − m]!
where [n]!=[1][2]..[n] and J+ |j, j >= 0, < j, j|j, j >= 1. Also J0 |j, m >=
m|j, m > and
p
J± | j, mi = [j ± m][j ± m + 1] | j, m ± 1i (11.71)
The Casimir invariant operator is given as

C2q = J− J+ + [J0 ][J0 + 1] (11.72)


whose eigenvalue is

C2q | j, mi =| [j][j + 1] | j, mi (11.73)


We wish to use this language for the deformed nuclei. As an extension of
the standard rotor model we formulate a q-rotor model based on the group
SUq (2). It is a system whose Hamiltonian is given as,

C2q
Hq = + E0 (11.74)
2I
Here I is the moment of inertia and E0 is the band head energy. For the
case of q being a phase we obtain,

[j][j + 1] sin(tj)sin(t(j + 1))


Ej = + E0 = + E0 (11.75)
2I 2I · sin2 t
For small t we can Taylor expand the series to look like:

[j(j + 1) − t2 (j(j + 1))2 /3 + 2t4 (j(j + 1))3 /45 − ..


Ej = E0 + (11.76)
2I
Now this is of the form:

Ej = E0 + A(j(j + 1)) + B(j(j + 1))2 + C(j(j + 1)3 + .. (11.77)

This is an empirical series used in fitting rotational bands in nuclei. It is


found that A,B,C.., possess alternating signs and that B/A is approximately
10−3 etc. Our series above has this characteristic for t approx .03 – a typical
value found in fitting rotational bands in nuclei. This is also similar to the
434 Group Theory in Particle, Nuclear, and Hadron Physics

VMI model. The above Taylor series was used [113] to obtain a reasonable
description of deformation and SD in nuclei.
However note that for higher j the above series does not converge fast
enough and hence, if we use it to fit SD with high j, the fits are likely to be
of poor quality. What we have done is to take the philosophy that we have
the exact nonlinear expression for Ej ( above) and this we can justifiably
use for any j (howsoever high). We therefore use the exact expression of Ej
(which has two parameters I and t) to fit the deformation and SD bands in
nuclei in Table 2 with experimental data from [114]. As expected we find that
we immediately make fits of superior quality than accomplished earlier [113].
For example for 192-Hg (2164 kev) we obtain a superior fit where the rms
deviation is 3.53 kev compared to the earlier fit [115] of rms deviation 8.26
kev. This is general conclusion that always the fits that we get are superior to
the one compared to the Taylor series expansion one [114].
The goodness of the fit is determined by the root-mean-square-deviation:

X Ej (cal)2 − Ej (exp)2 1/2


σ = [1/N [ ] (11.78)
Ej (exp)2
Now if this were all, then what we have achieved is basically a new for-
malism of the deformed quantum group SU (2)q which yields more convenient
and superior fits to the deformed and Superdeformed Bands in nuclei. But is
this all that this quantum group does for us?
There is however one aspect of the Superdeformed Bands which the stan-
dard classical SU(2) fails to account for. It turns out that a set of Superde-
formed Bands in odd-even nuclei of 152 151
66 Dy and 65 T b with integral-J and half-
integral-J, respectively, are exactly identical as shown in Table 11.3 (experi-
mental data from [116], [117]). There are several examples of identical bands
exhibiting an exact similarity between SD bands in integral-spin and adjoining
half-integral spin nuclei [116], [117].
Now in particle physics there is a currently popular idea of Supersymmetry
(SUSY) which advocates symmetry and identity between integral and half-
integral spin entities. Within the field of particle physics, however, as of the
beginning of 2016, there are no empirical indications of any support for this
concept. However here in case of the Superdeformed Bands we have several
examples of exact similarity and hence exact identity between the SD bands
of integral-spin and half-integral-spin nuclei. Amazingly this is a clear experi-
mental manifestation of Supersymmetry in nuclei. So Supersymmetry, which
is missing in particle physics empirical data, arises clearly and uniquely in a
different structure (of superdeformation) in the nucleus.
So how do we understand this new structure? Clearly the ordinary classi-
cal group SU(2) would fail here. But now we have this new language of the
quantum group. Could it be useful here? Now, we have fitted the identical
Superdeformed Bands in 152 Dy and 151 T b, and the results are depicted in Ta-
ble 11.3. Indeed the fit is remarkable. So the quantum group SU (2)q succeeds
Symmetry in Nuclei 435
TABLE 11.2: Experimental and
theoretical data for Superdeformed Bands in
194
Hg(2) compared with our results. σ=2.092
kev, t=0.0107 and 1/2I=532 kev for others
and σ=1.051 kev, t=0.011 and 1/2I=530 kev
for our result. Note J = J0 + 2.
J Expt[Mev] others Our fit
10 0.201 0.201 0.200
12 0.444 0.443 0.442
14 0.727 0.726 0.724
16 1.052 1.049 1.047
18 1.415 1.413 1.410
20 1.818 1.815 1.812
22 2.259 2.257 2.253
24 2.737 2.735 2.731
26 3.252 3.251 3.247
28 3.802 3.803 3.798
30 4.387 4.389 4.385
32 5.007 5.009 5.006
34 5.659 5.662 5.659
36 6.344 6.347 6.345 [116], [117].
38 7.062 7.061 7.061
40 7.809 7.805 7.807

in providing good and consistent fits to the identical bands in nuclei. Thus
the quantum group accomplishes something which the corresponding classical
group cannot dream of. However, at this stage this is merely a phenomenolog-
ical fit. We have to now determine as to which aspect of the quantum group
SU (2)q causes the property of Supersymmetry to exist within it. This discus-
sion is related to the ongoing debate as to whether the Lie algebra or the Lie
group is physically more fundamental [118]. As the quantum group is actually
a deformation of the corresponding Lie algebra with no corresponding simple
Lie group, clearly the Lie algebra has more relevance for the physical theories.
Thus in a Lie group having global structure, we can always proceed to the
local aspect by confining ourselves to the local Lie algebra. But as in the case
of the quantum [116], [117].algebra which is local, we have no corresponding
global aspect of the theory. Thus the SU (2)q quantum group is able to de-
scribe supersymmetry of the SD bands, which is an anathema for the standard
SU (2)I – isospin Lie group.
436 Group Theory in Particle, Nuclear, and Hadron Physics

TABLE 11.3: Supersymmetry in nucleus – fitting


of identical Superdformed Bands in the neighbouring
nuclei, 152 Dy and 151 T b∗ . σ is the rms deviation. For
152
Dy σ = 3.3x10−3 , t=1x10−5 and 1/2I=587 for
151
T b∗ Note J = J0 + 2.
152 151
Dy Tb
J Expt[kev] Our fit J Expt[kev] Our fit
26 602.2 598.46
28 647.2 645.4 27.5 647.0 640.31
30 692.2 692.34 29.5 692.0 687.69
32 737.5 739.27 31.5 738.0 735.06
34 783.5 786.21 33.5 783.0 782.42
36 829.2 883.15 35.5 828.0 892.76
38 876.1 880.09 37.5 876.0 877.09
40 923.1 927.03 39.5 992.0 921.41
42 970.0 973.96 41.5 970.0 971.72
44 1017.0 1020.9 43.5 1016.0 1019.0
46 1064.8 1067.84 43.5 1963.0 1066.28
48 1112.7 1114.78 47.5 1112.0 1113.53
50 1160.8 1161.72 49.5 1158.0 1160.77
52 1208.7 1208.65 51.5 1207.0 1207.99
54 1256.6 1255.59 53.5 1256.0 1255.19
56 1304.7 1302.53 55.5 1305.0 1302.37
58 1353.0 1349.47 57.5 1353.0 1349.53
60 1401.7 1396.41
62 1449.4 1443.34
Chapter 12
Quarks in Nuclei

12.1 The EMC Effect and the Nucleus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 437


12.2 Hidden Colour in Multiquarks in Nuclei . . . . . . . . . . . . . . . . . . . . . . . . . 438
12.3 Quarks in A=3 Nuclei . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 448
12.4 ∆ Excitations in the Nucleus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 449
12.5 M1 Strength in Nuclei . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 451
12.6 Gamow-Teller (GT) Strength in Nuclei . . . . . . . . . . . . . . . . . . . . . . . . . . 457
12.7 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 458

12.1 The EMC Effect and the Nucleus


As elucidated in the previous chapters, the framework conceptualising the
individual Z number of protons and N number of neutrons as confined in a
suitable potential-well essentially provides a very successful description of the
nucleus. In fact symmetry arguments, commencing with the isospin symme-
try SU (2)I , when taking (p,n) to provide the fundamental representation of
this group and after suitable expansion to the Generalized Pauli Exclusion
Principle, are very basic to a successful description of the nucleus.
In spite of the above, the proof that the three constituent quarks inside each
nucleon in a nucleus are not mere idle spectators, was provided unambiguously
by the so called “EMC effect” only as late as 1983.
Suppose we perform deep-inelastic muon scattering experiment on a single
nucleon and plot the structure function F2 (see Chapter 7) as a function of
x, the Bjorken variable. Next we perform the same experiment on a heav-
ier nucleus and plot the results of the new F2 as a function of the same x,
then what should one expect to find? Clearly, as the binding energy of nucle-
ons in a nucleus are several order of magnitude smaller than the momentum
transfers in deep inelastic scattering, then the above ratio should be unity.
Minor corrections due to the Fermi motion of nucleons in the nucleus may
be present. So whether we take a single nucleon or a complex nucleus, deep
inelastic scattering results should be identical.
Hence when the European Muon Collaboration  (EMC)  working at CERN,
published their results in 1983 [119], on A1 σ A / 21 σ D where A is the 56 F e
nucleus and D is the deuteron, the effect was to produce great surprise within
the scientific community. For they discovered an enhancement of this ratio for

437
438 Group Theory in Particle, Nuclear, and Hadron Physics

values above x ∼ 0.3 and suppression below it. This so-called EMC effect has
since then been confirmed by several other groups and for several other nuclei.
One popular feeling is that this phenomenon illustrates that the quarks in
individual nucleons are not mere passive spectators bound inside individual
nucleons, but partake in a highly active role in the nucleus. Thus the EMC
effect is clear empirical evidence in support of specific quark effects in nuclei.
Several models postulate that the size of bound nucleons is larger than
the free ones. In these frameworks, quarks in nuclei move in bags such as 6-,
9- and 12-quark bags, Valence quarks annihilate anti-quarks from the nuclear
sea. Here we shall study how the quarks in a single nucleon are affected by
the fact that there are other nucleons present in a nuclear medium. Also we
shall study specific situations where quarks provide their unique signatures in
a nucleus.

12.2 Hidden Colour in Multiquarks in Nuclei


We know that for a single baryon and a single meson in colour space
SU (3)C , we have, respectively,

3 ⊗ 3 ⊗ 3 = 1 ⊕ 8 ⊕ 8 ⊕ 10 (12.1)
3×3=1⊕8 (12.2)
The colour confinement hypothesis of QCD states that what we observe
as a baryon and a meson in the laboratory is the particle that corresponds
to the colour singlet part above. So what happens to the 8- and the 10-plet
parts? Though they play no direct role for single-hadrons but due to the fact
that a colour singlet state also arises in

8 ⊗ 8 = 1 ⊕ 8 ⊕ 8 ⊕ 10 ⊕ 10 ⊕ 27 (12.3)
So what happens to this singlet? Clearly when we take two bags of 3-
quarks each and if they overlap to form a bag of 6-quarks, then the six-quark
bag will be a colour singlet overall in two possible ways. The colour singlet
component of each 3-quarks bags will combine in the 6-quarks bags to give
colour singlet in a trivial manner. In addition the 8-plet component of colour
for each 3-quark system should merge into the 6-quark system as 8⊗8 to yield
a singlet as in the above equation. So,

|6q >Signlet = a {|1 >1 ⊗|1 >2 }Signlet ⊕ b {|8 >2 ⊗|8 >2 }Signlet (12.4)

where the number in state |x > (x = 1, 8) refers to the colour representation


and the number in subscript corresponds to the baryon number 1 or 2. Overall
Quarks in Nuclei 439

each product should be colour singlet individually. And also a2 + b2 = 1. Thus


this is the complete colour singlet state of a single composite 6-quark system.
Below we shall show that here a2 = 0.2 and b2 = 0.8. So the 6-quark
colour singlet state consists of two 3-quark bags in singlet states 20% of the
time each, and of two 3-quark bags in colour octet states 80% of the time
each. Together these two bags constitute a singlet.
What is the relationship of this 6-quarks colour singlet system with the
deuteron, the bound state of a lone proton and a lone neutron? As per the
conventional picture of the nucleus, the answer to the above question is – none!
Already the magnetic moment, the quadrupole moment, the root-mean-square
radius, the ground state binding energy, and various other properties of the
deuteron are very well understood, with the deformed deuteron made up only
of proton and neutron. So as per this picture there is no need to invoke quark
degrees of freedom.
But this perspective is in conflict with the EMC effect. Where is the catch?
The catch is that the N-N interaction – as required to fit the nuclei – generally
has a “hard core”, implying that for relative distances less than ∼ 0.4f m
between nucleons, the force becomes strongly repulsive.
This inconvenience is swept under the carpet by invoking a phenomenolog-
ical hard core. Clearly this hard core is outside the purview of nuclear physics,
requiring only nucleonic degree of freedom. Very often an ω- meson exchange
interaction is invoked [98] to explain this hard core. However, it is difficult
to visualize ∼ 0.6f m size ω-mesons being exchanged over ranges less than
∼ 0.4f m. Hence it is rather quarks which appear as the most logical source
of this hard core.
One may imagine the density distribution of the deuteron to be that of a
dumbbell – an entity comprising two fat balls at the ends joined by a narrow
tube. The centre of this nucleus is r = 0 where r is the relative distance
between the centre of masses of the two nucleons in the deuteron. Now as each
nucleon has a finite size ∼ 0.8f m, the r = 0 density distribution necessarily
corresponds to the two nucleons sitting on top of each other. Thus this is
clearly a 6-quark composite system as obtained in the equation above.
This can best be visualized in a hybrid model of deuteron. Take a scale
r0 ∼ 0.4f m (same as the size of the hard core in the N-N interaction potential).
For r > r0 in the deuteron the system consists of colour singlet neutron and
proton (no quark degrees of freedom here), and for r < r0 the system is that
of 6-quarks as given above. Now for r < r0 as given above, 20% of the 6-quark
colour singlet is constitute in terms of individual 3-quark colour singlet states,
which free nucleons anyway are. Meanwhile, for 80% of the time these reside in
8-coloured components which free nucleons are never in. Hence this part of the
wave function can never exist as free states and are condemned to be available
only in the inner core (r < r0 ) region of the deuteron. Consequently, this 80%
component of the wave function of 6-q are hidden inside the deuteron. Hence
this is denominated the hidden colour part of the 6-q composite states.
Now clearly the system, i.e., the deuteron, predominantly exists in indi-
440 Group Theory in Particle, Nuclear, and Hadron Physics

vidual colour singlet states of a proton and a neutron. Hence the large hidden
colour component should act as an effective repulsion mechanism between
the colour singlet proton and neutron, leading to the formation of the well-
documented dumbbell-shaped density distribution. The hidden colour part
would also be the source of the hard core between nucleons in the nucleus.
This repulsive role of the hidden colour components in the multiquark
system is a generic property of the nuclear realm. Now we already saw in the
previous chapter that the SU (4)SI group of Wigner cannot be entertained as
a serious model. We pointed out that the nuclear potential hard core at short
range is what is required to explain nuclear saturation ([99], p.82).
Now here we show that this inter-nucleon short-range repulsion mechanism
is a natural consequence of the hidden colour components of the multiquark
model. In fact, this hard core is to be viewed as a significant unambiguous
signature of the quark effect in nuclei.
Now let us move our focus to the next heavier nuclei, A = 3, 3 He and 3 H
and A = 4,4 He. In particular, let us analyse the situation at r = 0, namely
at a zero relative distance between nucleons, which is essentially the centre of
these nuclei. In this case, the 9-quark configuration for A = 3 nuclei and the
12-quark configuration for A = 4 nuclei should be significant.
So what are the hidden colour components in 9-quark and 12-quark sys-
tems? Obtaining the hidden colour component for 6-quark entities was labou-
rious as it is [52], and deriving the hidden colour component for the 9-quarks
and the 12-quarks is indeed a rather intensive group theoretical exercise [120].
However, due to the significance of the final result, the mathematics shall be
discussed below. At the outset, let us mention that the ultimate conclusion is
that the hidden colour components in 9-quark and 12-quark systems are gi-
gantic in nature, representing an overwhelming 97.6% and 99.8% of the total,
respectively.
It is to be noted the rms-radii of 3 He, 3 H and 4 He are quite compact,
numbering 1.88 fm, 1.70 fm, and 1.64 fm, respectively. Now, the fact is that
each nucleon has a size of ∼ 1f m, which is very clearly off the centre of the
mass of these nuclei. Hence, at r = 0 in A = 3 nuclei the 9-quarks configuration
would be dominant. Likewise, in A = 4 nuclei the 12-quarks configuration
would be dominant.
Consequently, we are utilizing a hybrid picture wherein for r > r0 the
configuration is a three-colour singlet system and for r < r0 it is a 9-quark
system. As these r < r0 are predominantly hidden colour in nature, there
would exist an effective repulsion pushing out these nucleons away from the
centre. Thus as per the quark model picture there would be a “hole” in the
central density distribution of these nuclei. Indeed this prediction is precisely
what is observed in exhaustive experiments [121].
It should be remarked that the density distribution of nuclei is well stud-
ied through electron scattering experiments [121]. In specialist literature it is
commonplace to plot this density distribution as a function of r, the relative
distance between nucleons in nuclei. The r = 0 point is truly the centre of the
Quarks in Nuclei 441

i.o

0.8

Experiment
0.6 Theory
Pd

0.4

o.:

0
0 0.5 1.0 1.5 2.O 2.5 3.0

r(fm

FIGURE 12.1: Experimentally determined central hole in the charge density


distribution of A=3 and A=4 nuclei and the standard theoretical expectation

nucleus. Note that this characteristic is also a genuine ground state property
of the nucleus. The experimental result [121] and the theoretical expectation
based on the canonical understanding of the nuclear density distribution is
schematically shown in Figure 12.1.
The hole is a puzzle within our conventional understanding of the nuclear
density distribution. Thus to be able to explain the short-range repulsion be-
tween two nucleons, we have to go beyond conventional picture of the nucleus
having only protons and neutrons as its sole degrees of freedom. We shall
show that to be able clarify the hole at the centre of the density distribution
of A=3 and A=4 nuclei, the hidden colour components of the corresponding
multiquark system play a decisive role.
Having discussed the physics arising from hidden colour components in
multiquark systems, we now discuss the group theoretical aspects of obtaining
the proper wave function in the colour space. The hidden colour components
were strenuous to obtain for the 6-quark system [52] and were more intricate
still for the 9-quark and 12-quark systems.
The difficulty in group theoretical calculation lies in the fact that, as we
saw in the product of representations in Ch-5, this entails calculation of the
Coefficient of Fractional Parentage (CFP) of large SU (N ) group decomposi-
tion and that of the Isoscalar Factors (ISF) of a large permutation group S(n)
decomposition.
Fortunately it emerges that there exists a direct relation between the CFP
of the unitary group decomposition SU (mn) ⊃ SU (m) ⊗ SU (n) and the ISF
442 Group Theory in Particle, Nuclear, and Hadron Physics

of the permutation group chain S (f1 + f2 ) ⊃ S (f1 ) ⊗ (f2 ) ; that is that these
two are identical. Therefore these CFPs are independent of m and n and are
given by [43],
{ν}τ,β{σ}θ{µ}φ
C{ν 0 }β 0 σ0 µ0 ,{ν 00 }β 00 σ00 µ00 (12.5)
where {σ} , {µ}and {ν} are partition labels for the irreducible representations
of SU (m), SU (n) and SU (mn), respectively; β 0 , β 00 and β are inner multiplic-
ity labels, for example, β 0 = 1, 2... (σµ0 ν 0 ), etc. And θ, φ and τ are the outer
multiplicity labels, for example, θ = 1, 2, ... (σ 0 σ 00 σ), etc.
These CFPs are quite complicated objects. Fortunately for us, these sim-
plify for some special cases, When τ does not play any significant role and when
antisymmetric representations like [ν 0 ] = 1f1 ,

we are dealing with totally
[ν 00 ] = 1f2 ; [ν] = 1f , of S (f1 + f2 ) ⊃ S (f1 ) ⊗ (f2 ) decomposition, the C
 

simplifies to

  21
[1f ],[σ]θ[µ]φ hσ0 hσ00
C 1f1 ,[σ0 ][µ0 ], 1f2 ,[σ00 ][µ00 ] = δµσ̃ δµ0 σ̃0 δµ00 σ̃00 δθδ (12.6)
[ ] [ ] hσ
Now [σ] , [µ] , [θ] are the partition labels for the irreducible representation
of the groups SU (m), SU (n) and SU (mn). δµσ̃ means µ = σ̃. Because of the
deltas it reduces to calculating for [σ] i.e the SU (m) cases only. Here hσ , hσ0
and hσ00 are dimensions of the irreducible representation of the permutation
group S(f ), S(f1 ) and S(f2 ), respectively. Note that the conjugate represen-
tation is specified by a twiddle in the delta functions.
Now with the SU (2)I isospin group and with (p, n) forming the fundamen-
tal representation of it (and thus p and n are indistinguishable fermions here)
and if we expand to SU (4)IS ⊃ SU (2)I ⊗ SU (2)S with
 ↑
p
 p↓ 
 ↑
n  (12.7)

n
This provides the fundamental representation of the product group
SU (4)SI . It means that p↑ , p↓ , n↑ , n↓ reside as degenerate entities in the


lowest ground state orbital 0S in the confining potential. So up to four of


these states can be found at the same point as per the Pauli Exclusion Prin-
ciple. Not to forget that had it been distinguishable protons and neutrons,
Pauli Exclusion Principle would allow only two states p↑ , p↓ and

then the
n↑ , n↓ in a bound 0S single orbital each.


Now let us consider only (v, d) quark in SU (2)F subgroup of SU (3)F


group. In nuclear physics we assume no strange quark exists in low energy
excitations. Now in SU (4)F S ⊃ SU (2)F ⊗ SU (2)S as a good symmetry, we
have u↑ , u↓ , d↑ , d↓ degeneracy in one orbital. If we include colour SU (3)c
and enlarge the group to
Quarks in Nuclei 443

SU (12)SF C ⊃ SU (4)SF ⊗ SU (3)C ⊃ SU (2)5 ⊗ SU (2)F ⊗ SU (3)C (12.8)


and if this is a good group, then we can have up to 12 quarks with spin-colour
degrees of freedom in the ground state. Now for a single nucleon we throw out
the colour singlet representation as providing us with the antisymmetry for the
three fermionic quarks and thus the SU (4)SF state is symmetric to describe N
and ∆ degrees of freedom. But now for multiquark systems, as higher colour
representations (such as, for example, 8 ⊗ 8 in the 6-quarks case) yield colour
singlet states, the relevant group to consider is not just SU (4)⊗SU (3) but the
more expansive group SU (12)SF C . Hence we require CFP of SU (12)SF C ⊃
SU (4)SF ⊗ SU (3)C decomposition for multiquark states.
For three-flavour quarks in the quark model with the group SU (6)SF ⊗
SU (3)C , the ground states are given as

⊗ (12.9)

(56)SU (6)SF ⊗ 1SU (3)C (12.10)


We can consider this as fully antisymmetric state of SU (18)SF C as de-
composed with respect to subgroup SU (6)SF ⊗ SU (3)C . Note that the Young
Diagrams in the two subgroup are conjugates of each other.
So we carry this logic further for the 6-quark singlet in the group
SU (12)SF C ⊃ SU (4)SF ⊗ SU (3)C . We find the fully antisymmetric state
only from the combination of the conjugate diagrams in the two subgroup as

⊗ (12.11)

(56)SU (4)SF ⊗ 1SU (3)c (12.12)


Now if we take 4 as the fundamental representation SU (4)SF and describe
the product for 6-quark as 4 ⊗ 4 ⊗ 4 ⊗ 4 ⊗ 4 ⊗ 4, then one finds that the
above Young Diagrams of SU (4)SF occurs five times independently. Also, if
we take 3 as the fundamental representation of SU (3)c , then in the product of
3 ⊗ 3 ⊗ 3 ⊗ 3 ⊗ 3 ⊗ 3 the above Young Diagram of SU (3)c also occurs five times.
This signifies that the fully antisymmetric state of the 6-quarks for the group
SU (4)SF ⊗SU (3)C is not unique. Thus we cannot build a good antisymmetric
state of the 6-quarks within this group structure. This obstacle arises for a
multiquark system of 6-quarks. Such a quandary did not arise for the 3-quarks
system. In fact it was for this group that we determined a good antisymmetric
state for the 3-quark system. So the group SU (4)SF ⊗ SU (3)C which yielded
a good antisymmetric state for 3-quarks does not provide us with a good
444 Group Theory in Particle, Nuclear, and Hadron Physics

antisymmetric state for the 6-quark system. This is a new constraint arising
in the multiquark system in contrast to the bound system of 3-quarks.

————————————————–
Problem 12.1: Confirm that the Kroenecker product in SU (4)SF case
and the SU (3)c cases, the (56)SU (4)SF and 1SU (3)c Young Diagrams occur
five independent times each.
————————————————–

Thus there is no unique good antisymmetric state of 6-quarks for the


group SU (4)SF ⊗ SU (3)c . Now this ambiguity due to the redundancy has to
be eliminated to obtain a proper antisymmetric state for 6-quarks. This is
accomplished by embedding the group SU (4)SF ⊗ SU (3)c in a larger group
and consequently applying antisymmetrisation techniques. In fact, this task
can be achieved for the larger group SU (12)SF C under the decomposition
SU (12)SF C ⊃ SU (4)SF ⊗ SU (3)c
In the group SU (12)  C we take the product of six-quarks as 12 ⊗ 12 ⊗
 SF
12 ⊗ 12 ⊗ 12 ⊗ 12 i.e. 16 in SU (12). We establish the dimension of the fully
anti symmetry states as

(12.13)

 6 12 × 11 × 10 × 9 × 8 × 7
1 A= = 924 (12.14)
6×5×4×3×2×1
Now this unique
 and completely antisymmetric SU (12) representation of
dimension 16 = 924 may be reduced with respect to the subgroup SU (4) ×
SU (3) as,

924 = (50, 1) ⊕ (64, 8) ⊕ (6, 27) ⊕ (10, 10) ⊕ 10, 10 (12.15)

————————————————–
Problem 12.2: Confirm that the above reduction in Equation 12.15 is the
proper reduction with respect to the states and the groups under discussion.
————————————————–

Now we see that the antisymmetric representation (50, 1) (as in Equation


12.11 is contained only once in the 924-dimensional antisymmetric representa-
tion. This shows that the 6-quarks states are consistently classified as elements
Quarks in Nuclei 445

of the (50, 1) representation of SU (4)SF ⊗ SU (3)C of the 924-dimensional


anti-symmetric representation of SU (12)SF C . Therefore this is the good and
unique antisymmetric state of the 6-quarks.
But it requires that the group SU (12)SF C should exist and that the above
decomposition SU (12)SF C ⊃ SU (4)SF ⊗ SU (3)C be valid. Note the fine dis-
tinction that we are making when we use the word “exist” in the previous sen-
tence. Obviously the larger group always “exists” mathematically. But physi-
cally it may not “exist” at all. Hence, it is pertinent to question whether it is
justified to provide the fundamental representation of SU (12)SF C , which will
yield the fully antisymmetric state for a ground state as below.

u↑R
u↓R
u↑B
u↓B
u↑G
u↓G
d↑R
d↓R
d↑B
d↓B
d↑G
d↓G (12.16)
Just as the proton and neutron are treated as indistinguishable and identi-
cal particles in the isospin group SU (2)I , so are all these 12-states of the quark
equivalent for the group SU (12)SF C . Is this justified? The spin-SU (2)S group
is a good and fundamental exact structure in physics; the SU (3)c of QCD is
an exact symmetry; and given the fact that the isospin-SU (2)I symmetry is
a pretty good approximate symmetry; we may justifiably conclude that the
above fundamental representation of the larger group SU (12)SF C should be a
reasonably good approximate symmetry of hadrons. The same conclusion may
be arrived at with three light quarks in (u, d, s) with three-flavour symmetry
and the large group structure as, SU (18)SF C ⊃ SU (6)SF ⊗ SU (3)c .
But we shall find that the same cannot be said when heavier mass
quarks such as the c-,b- and t-quarks are included. This is because of the
fact that even for the group [SU (n)F n = 4, 5, 6] the symmetry group
is very strongly broken. For example, it would be dubious to utilise the
fundamental representation of the group when the c-quark is included as
SU (24)SF C ⊃ SU (8)SF ⊗ SU (3)c . So with heavy quarks you may build a
genuine 3-quark antisymmetric state, but none for the multiquarks like the
6-quarks and others.
This clearly has implications for the ongoing search in the laboratory of
multiquark states. We saw one such case for the pentaquarks in Chapter 7.
446 Group Theory in Particle, Nuclear, and Hadron Physics

Now for a 3-quarks system the states in SU (12)SF C are,

⊗ ⊗ = ⊕ ⊕ ⊕
12 ⊗ 12 ⊗ 12 = 364 ⊕ 572 ⊕ 572 ⊕ 220 (12.17)
(check for the last Young Diagram dimension is (12 × 11 × 10) / (3 × 2 × 1) = 220).
Let us look at the 220-dimensional antisymmetric representation in
SU (12)SF C . With respect to SU (4)SF ⊗ SU (3)c this reduces as follows

= ( , ) ⊕ ( , ) ⊕ ( , (12.18)

220 = (20, 1) ⊕ (200 , 8) + 4, 10



(12.19)
Now putting these three-quark antisymmetric states together to obtain
appropriate 6-quark states,

⊗ = ⊕ ⊕ ⊕ (12.20)

220 ⊗ 220 = 924 ⊕ 8580 ⊕ 23166 ⊕ 15730 (12.21)


6

Thus the 924-dimensional representation of 1 in SU (12)SF C occurs only
once in the above product.
Next note that from Equation 12.19,

⊗ 1 1 1 = 1 1 1 ⊕ others
20 ⊗ 20 = 50 ⊕ others
and

1 1 1
⊗ 2 = 1 2 ⊕ others

200 ⊗ 200 = 50 ⊕ others (12.22)


Quarks in Nuclei 447

So
(20, 1) ⊗ (20, 1) = (50 ⊕ · · · , 1) (12.23)
and
(200 , 8) ⊗ (200 , 8) = (50 ⊕ · · · 1 ⊕ · · · ) (12.24)
So (50, 1) in Equation 12.11 have two contributions as above, i.e., a colour
singlet 1 × 1 and the other 8 × 8 terms as in Equation 12.2 and 12.3. Let us
calculate these for CFP in Equation 12.6. Now recall that the dimension of
an irreducible representation of a Young Diagram in Sn is
n!
h[p] = (12.25)
hook − length − products
Hence
3!
[3] = : h[3] = =1 (12.26)
1×2×3

3!
[21] = : h[21] = =2 (12.27)
1×3×1

6!
[33] = : h[33] = =5 (12.28)
2×3×4×1×2×3
So for the colour singlet case in Equation 12.4 and 12.23,
  12 r
[16 ][33][33˜] h[3] h[3] 1
C[13 ][3] 3̃ ,[13 ][3] 3̃ = = (12.29)
[] [] h[33] 5
And for 8 × 8 case Equation 12.3 using Equation 12.24 becomes,
  12 r
[16 ][33][33
˜] h[21] h[21] 4
C[13 ][21] 21
˜ = = (12.30)
[ ],[13 ][21][21] h[33] 5
Whence finally the 6-quarks colour singlet states in Equation 12.4 are

r r
1 4
|6q >Singlet = {|1 >1 ⊗|1 >2 } + {|8 >1 ⊗|8 >2 }Singlet (12.31)
5 5
Thus the 6-quarks colour singlet state has 80% hidden colour components.
Through some similar and lengthy algebra one finds [120] the colour singlet
components in the 9-quarks and 12-quarks system as given in Table 12.1.
The rest are theqhidden colour parts. So for a 9-quarks system the hidden
41
colour represents 42 or 97.6% of the total 9-quarks colour singlet state
[120]. The hidden colour concept is also discussed in the case of pentaquarks
in Chapter 6.
448 Group Theory in Particle, Nuclear, and Hadron Physics
TABLE 12.1: Colour
singlet components in 6-, 9-
and 12-quark systems - the
rest are hidden colour

6-q 9-q 12-q

q q q
1 1 1
5 42 462

————————————————–
Problem 12.3: Obtain just the colour singlet part of the 9-quark colour
singlet wave function as specified in Table 12.1.
————————————————–

12.3 Quarks in A=3 Nuclei


3
In 2 Hethe two protons being identical fermions, must be antisymmetric in
spin space. Hence S = 0 for this pair and thus the spin of 3 He arises entirely
from the odd neutron. So we take
↑ S=0
n↑
3
He = (pp) (12.32)
and hence the (pp) pair does not contribute to its magnetic moment. So the
magnetic moment of 3 He is entirely due to the odd neutron. Similarly the
magnetic moment of 3 H is due to its odd proton.
3

µ He = µ (n)
µ 3 H = µ (p)

(12.33)
Now for the deformed nucleon we demonstrate that the magnetic moment
of a free nucleon is modified by the factor (1 − PD (N )) where PD (N ) is the
deformation of the nucleon. Furthermore, let us postulate that the nucleon
may be modified by the nuclear medium. Hence, for nucleons inside A = 3
nuclei we assume

3 2
He = − 1 − PD 3 He µq
 
µ
3
µ 3 H = 1 − PD 3 H µq
  
(12.34)
Quarks in Nuclei 449

where PD (N ) corresponds to the deformation of the odd nucleon for a partic-


ular nucleus. Taking the ratios
   
µ 3 He 1 − PD 3 He µ 3H 1 − PD 3 H
= ; = (12.35)
µ(n) 1 − PD (N ) µ(p) 1 − PD (N )

we take the experimental values (in nuclear magneto) µ 3 He = −2.12,
µ(n) = −1.91 ; µ 3 H = 2.98, µ(p) = 2.79 and using PD (N ) = 14 (which


gives good fits to all the experimental values) we obtain PD 3 He = 0.168


and µ 3 H = 0.199. Next

   
δµ 3 He µ 3 He − µ(n) 2 PD (N ) − PD 3 He
= = − (12.36)
δµ (3 H) µ (3 H) − µ(p) 3 [PD (N ) − PD (3 H)]

We obtain -1.08 for the above ratio with values calculated above. This fits
very nicely with an experimental value of -1.1 .
One of the expected generic properties of the nucleon in the nuclear
medium is an increase of the effective confinement or bag size. For a deformed
baryon this implies we visualize a larger surface area in the flat region rather
than the edges as say in a pumpkin. Let us assume that uniform pull exists
on the surface of a nucleon in a nucleus (to increase its size as per the above
scenario). In that case there is more pull exerted on the flat-deformed surface
compared to the edges and thus the overall anamorphosis of the nucleon with
decrease in deformation in the nuclear medium. And this is indeed what is
obtained for 3 He and 3 H as PD (A) < PD (N ).
We know that for 3 He the binding energy is -7.77 MeV and r(rms) =1.88
fm while for 3 H the binding energy is -8.48 MeV and r(rms) = 1.70 fm. Hence
3
He is a more loosely bound A = 3 system than 3 H. As there is a larger space
available for a nucleon to expand in 3 He than than in 3 H, the effect in PD (A)
is naturally larger for 3 He than for 3 H [122].

12.4 ∆ Excitations in the Nucleus


Let us assume that our nucleus consist of a number of nucleons antisym-
metrised with respect to each other. Let each nucleon consist of 3-quarks with
proper symmetry. Hence the nuclear ground states is
1
|0 >= √ A ΠA
i=1 | (qqq)i > (12.37)
A!
where |qqq >i,i=1,2....A is the quark model nucleon wave function. We assume
that the specific nuclear correlations modify this quark model state. This
450 Group Theory in Particle, Nuclear, and Hadron Physics

model is motivated by and justified by the EMC effect where specific nuclear
medium effects modify individual nucleon properties.
Now we have an operator like the magnetic moment operator M which acts
on each nucleon and also on quarks within each nucleon. Now this magnetic
moment operator, at the quark level, can excite only the N- and the ∆- degrees
of freedom. Now we calculate the total amount of strength excited by the M
operator. It is given as

X A
X A
X
S= | < β| M (i)|0 > |2 − | < 0| M (i)|0 > |2 (12.38)
β i=1 i=1

Where |0 > is the ground state of the nucleus of operator M acting on


it. This can transform to the same state |0 > or to excited states |β >. The
second term in Equation 12.36, when subtracted (as given above) produces
the genuine excitation strength due to the operator M. Now,
X
|β >< β| = 1 (12.39)
β

for the complete set of nuclear final states. So

X X
S = | < 0| M (i)M (j)|0 > |2 − | < 0| M (i)|0 > |2 (12.40)
i,j i=1

The second term is


X
| < 0| M (i)|0 > |2 = (T rM )2 (12.41)
i=1

while the first one is,


X X X
| < 0| M (i)M (j)|0 > |2 = | < 0| M (i)M (i)|0 > |2 +| < 0| M (i)M (j)|0 > |2
i,j i i6=j
(12.42)

0
X
= T r(M M 0 ) + (< ψk (i)ψt (j)|M (i)M (j) |ψk (i)ψt (j) > (12.43)
k,t

0
− < ψk (i)ψt (j)|M (i)M (j) |ψt (i)ψk (j) >) (12.44)

X
= T r(M M 0 )+ (< ψk |M |ψk >< ψt |M 0 |ψt > − < ψk |M |ψt >< ψt |M 0 |ψk >
k,t
(12.45)

= T r(M M 0 ) + T r(M ) · T r(M 0 ) − T r(M M 0 ) (12.46)


Quarks in Nuclei 451

Putting it all together,

S = T r(M M 0 ) + T r(M ) · T r(M 0 ) − T r(M M 0 ) − (T rM )2 (12.47)

Note that the bar over the first term indicates the sum over the same nu-
cleon, while in the third term the intermediate states are another intermediate
N or ∆ excited states. Note the difference in M and M 0 .

12.5 M1 Strength in Nuclei


For the magnetic moment we used the z-component of the magnetic mo-
ment operator because that is the manner in which the ground state magnetic
moment of a nucleon or a nucleus is obtained. But for the M1 strength in nu-
clei the magnetic moment M1 operator would act on all the spin components.
And hence for our model we here define the M1 operator as M and M 0 in
Equation 12.53.
A X
X A X
M= ( σµ (i)Q(i)) (12.48)
a=1 i=1 µ

A X
X A X
M0 = ( (−1)µ σ−µ (i)Q(i)) (12.49)
a=1 i=1 µ

Here index a sums over A-nucleons and for each nucleon the quark model
state is summed over the index i = 1, 2, 3 where
σx + iσy + σx − iσy
σ+1 = − √ ; σ+1 =− √ = −σ−1
2 2
σx − iσy + σx + iσy
σ−1 = √ ; σ−1 = √ = −σ+1
2 2
√ √ µ
= 2σ− and σ+1 = − 2σ+ ; σµ+ = (−1) σ−µ

σ−1 (12.50)
Now let us evaluate the different trace terms in Equation 12.47.
First
A
X XX
T rM = < ψk | σµ (i)Q(i)|ψk >
k=1 i µ

X X
= < ψk | (σ+1 (i)Q(i) + σ−1 (i)Q(i) + σz (i)Q(i)) |ψk > (12.51)
k i
452 Group Theory in Particle, Nuclear, and Hadron Physics

Now |ψk > are all the occupied orbitals of the nucleus in the ground state.
Since σ+1 and σ−1 are proportional to the raising and the lowering operators
so in the trace these give zero. Thus

X 3
X
T rM = < ψk | σz (i)Q(i)|ψk >= T r(M ) (12.52)
k i=1

This is also equal to T r(M 0 ) and so,

T r(M ).T r(M 0 ) = (T rM )2


Hence the 2nd and 4th terms in Equation 12.47 cancel. Thus (T r(M 0 M ) =
T r(M M 1 ) etc.).

S = T r(M 0 M ) − T r(M 0 M ) (12.53)

3 X 3 X
X X 0 X
T r(M 0 M ) = < ψt |[ (−1)µ σ−µ0 (i)Q(i)][ σµ (j)Q(j)]|ψt >
t i=1 µ0 j µ

X X XX 0
= < ψt | [ (−1)µ σ−µ0 (i)σµ (i)]Q(i)Q(i)|ψt >
t i µ0 µ
X XX 0
+ < ψt | [ (−1)µ σ−µ0 (i)σµ (j)]Q(i)Q(j)|ψt > (12.54)
t i6=j µµ0

We get
XX 0 X X 0
(−1)µ σ−µ0 (i)σµ (j) = (−1)µ σ−µ (i)σµ (i) + (−1)µ σ−µ (i)σµ (i)
µ µ0 µ µ6=µ0


= 3 − 2 2σz (i) = 3 − 2 (σ+1 (i) − σ−1 (i)) (12.55)
Now the first term in Equation 12.54 substituted into the second term in
Equation 12.52 yields zero.
3
X X √
< ψt | [−2 2(σ+ (i) − σ− (i))]t|ψt >= 0
k i=2

So the first terms is

X 3
X
=3 < ψt | Q(i)2 |ψt > (12.56)
t i=1

————————————————– P3
Problem 12.4: Show that t < ψt | i=1 Q(i)2 |ψt >= Z + 32 N
P
Quarks in Nuclei 453

————————————————–

Thus the first term is equal to 3Z + 2N . Next the second term in Equation
12.54,
X XX 0
< ψt | [ (−1)µ σ−µ0 (i)σµ (j)]Q(i)Q(j)|ψt >
t i6=j µµ0
X X
= < ψt | [−σ−1 (i)σ+1 (j) − σ+1 (i)σ−1 (j) + σz (i)σz (j)]Q(i)Q(j)|ψt >
t i6=j
(12.57)
Here the third term is

8 2
Σ < ψt Σ σz (i)σz (j)Q(i)Q(j) ψt >= Z + N (12.58)
t i6=j 9 3

Next
X X X X
< ψt | −σ+1 (i)σ−1 (j)Q(i)Q(j)|ψt >= 2 < ψt | σ+ (i)σ− (j)Q(i)Q(j)|ψt >
t i6=j t i6=j

3X ρ ρ λ λ ρ ρ λ λ
= 2· < φ χ +φ χ |σ+ (3)σ− (1)Q(3)Q(1)+σ+ (3)σ− (2)Q(3)Q(2)|φ χ +φ χ > (12.59)
2 t

Now
1
< χ↑λ |σ+ (3)σ− (1)|χ↑λ >= − =< χ↑λ |σ+ (3)σ− (2)|χ↑λ >
3

1
< χ↑ρ |σ+ (3)σ− (1)|χ↑λ >= √ = − < χ↑ρ |σ+ (3)σ− (2)|χ↑λ >
3

< χ↑ρ |σ+ (3)σ− (1)|χ↑ρ >= 0 =< χ↑ρ |σ+ (3)σ− (2)|χ↑ρ >

< χ↑λ |σ+ (3)σ− (1)|χ↑ρ >= 0 =< χ↑λ |σ+ (3)σ− (2)|χ↑ρ > (12.60)
For proton,
1
< φλ |Q(3)Q(1)| φλ >= − =< φλ |Q(3)Q(2)| φλ >
9
3
< φρ |Q(3)Q(1)| φλ >= √ =< φρ |Q(3)Q(2)| φλ >
2 12
1
< φρ |Q(3)Q(1)| φρ >= − =< φρ |Q(3)Q(2)| φρ >
9
2
< φρ |Q(3)Q(1)| φρ >= √ = − < φχ |Q(3)Q(2)| φρ >
3 12
454 Group Theory in Particle, Nuclear, and Hadron Physics

For neutron,
1
< φλ |Q(3)Q(1)| φλ >= − =< φλ |Q(3)Q(2)| φλ >
6
1
< φρ |Q(3)Q(1)| φλ >= √ =< φρ |Q(3)Q(2)| φλ >
3 12
1
< φρ |Q(3)Q(1)| φρ >= − =< φρ |Q(3)Q(2)| φρ >
6
1
< φρ |Q(3)Q(1)| φρ >= √ = − < φχ |Q(3)Q(2)| φρ >
3 12
Which in Equation 12.59 gives 3 · 98 Z + 23 N . Thus,


   
2 8 2 17
T r(M 0 M ) = 3 · Z + N + 3 · Z+ N = Z + 4N (12.61)
3 9 3 3
Next
X X
T r(M 0 M ) = < ψt |M 0 M | ψt >= < ψt |M 0 | ψk >< ψk |M | ψt >
t t,k

In this we have to evaluate,


XX
< ψt | σµ (i)Q(i)|ψk >
i µ
X √ √
=< ψt | [− 2σ+ (i)Q(i) + 2σ− (i)Q(i) + σ3 (i)Q(i)]|ψk >
i
P
For neutrons in nuclei this is (summation i is understood),
√ √
↓ ↑ ↑ ↑ ↑ ↓ ↓ ↓
=< n + 2σ− Q n > + < n |σz Q| n > + < n − 2σ+ Q n > + < n |σz Q| n >
(12.62)
µ
Noting that the other term has (−1) and thus for both p and n,
√ √
T r(M 0 M ) =< n↑ + 2σ+ Q n↓ >< n↓ 2σ− Q n↑ >

√ √
+ < n↑ |σz Q| n↑ >< n↑ |σz Q| n↑ > + < n↓ − 2σ− Q n↑ >< n↑ − 2σ+ Q n↓ >

+ < n↓ |σz Q| n↓ >< n↓ |σz Q| n↓ > + ( Same terms n replaced with p )


As
X X
< p↑ | Q(i)σz (i)|p↑ >= 1 =< p↓ | Q(i)σz (i)|p↓ > (12.63)
i

and
X 2 X
< n↓ | Q(i)σz (i)|n↓ >= − =< n↑ | Q(i)σz (i)kn↑ > (12.64)
3
Quarks in Nuclei 455

So the corresponding terms above for the nucleus give,


4
Z+ N (12.65)
9
Note that N ↑ + N ↓ = 2N and Z ↑ + Z ↓ = 2Z.
We also find
√ √
< p↑ 2σ+ Q p↓ >< p↓ + 2σ− Q p↑ >

= 2 < p↑ |σ+ Q| p↓ >< p↓ |σ− Q| p↑ >= 2Z (12.66)


and
√ √ 8
< n↑ + 2σ+ Q n↓ >< n↓ + 2σ− Q n↑ >= N (12.67)

9
Adding Equations 12.65, 12.66, and 12.67 we get,
     
4 8 4
T r(M 0 M ) = Z + N + 2Z + N = 3 Z + N (12.68)
9 9 9
Putting Equations 12.64 and 12.68 in Equation 12.53 the total strength is,
 
17 4
S= Z + 4N − 2Z − N
3 3

Call it total for (N + ∆)


8
S N +4 = (Z + N ) (12.69)
3
This is the total strength which signifies the fact that, due to the inclusion
of the quark model, it includes both the ground states of the N-kind and the
excited states of the ∆-kind in it.
Next we wish to calculate as to how much strength exists in the ∆-sector.
This is easier to derive, as ∆ is not Pauli blocked in the nucleus and thus each
neutron and proton can be excited to ∆. Thus,

X X X X
S∆ = < ∆| σµ (i)Q(i)|n > + < ∆| σµ (i)Q(i)|p > (12.70)
n i,µ p i,µ

where the first sum is over the neutrons and the second one over the protons
in the nucleus. Now for a pair of neutrons n↑ and n↓ this requires
X X
< ∆| σµ (i)Q(i)|n >=< ∆| σ+1 (i)Q(i)|n >
µ i
√ X √ X
=< ∆0− 1 , 3 | − 2 σ+ (i)Q(i)|n↑ > + < ∆0− 1 ,− 1 | 2 σ+ (i)Q(i)|n↑ >
2 2 2 2
i i
456 Group Theory in Particle, Nuclear, and Hadron Physics
X √ X
+ < ∆0− 1 , 1 | σj (i)Q(i)|n↑ > + < ∆0− 1 , 1 | − 2 σ+ (i)Q(i)|n↓ >
2 2 2 2
i
√ X √
+ < ∆0− 1 ,− 3 | 2 σ− (i)Q(i)|n↓ > + < ∆0− 1 ,− 1 | − 2Σσz (i)Q(i)|n↓ >
2 2 2 2 i

We get
8 ↑ 8 ↓ 8
n + n → N (12.71)
3 3 3
By symmetry we obtain the same result for the proton and thus
8
S4 = (N + Z) (12.72)
3
Now for the N = Z case for each n↑ orbital there exists a n↓ orbital and
for each p↑ orbital there is a p↓ orbital. But for the odd neutron n↑ or p↑ there
is no n↓ or p↓ orbital. So, for example, the last n↑ ,
√ √
< n↑ | + 2σ+ Q|n↓ >< n↓ | 2σ− Q|n↑
and only the σz Q term contributes. Thus the terms in Equation 12.66 and
Equation 12.67 are zero and Equation 12.65’s terms contribute, which for
T r(M 0 M ) is 49 and for T r(M 1 M ) = 4. Similarly for odd proton also. Hence
N
finally (note S = S N +4 − S 4 ),
8 0 8 0 14 32
S N +∆ = Z + N + δ( Z−1 ,I ) + δ( N −1 ,I ) (12.73)
3 3 3 2 9 2

8 0 8 0 8 8
S∆ = Z + N + δ( Z−1 ,I ) + δ( N −1 ,I ) (12.74)
3 3 3 2 3 2

8
S N = 2δ( Z−1 ,I ) + δ( N −1 ,I ) (12.75)
2 9 2

0 0
where N , Z are neutron and proton numbers for the corresponding even
core.
Now these results hold as per the spherical nucleon quark model. The
ground state M1 strength is zero for even-even nuclei. Experimentally for 28 Si
and 48 Ca the values are ∼ 7µ2N and ∼ 6µ2N . These are small in magnitude but
not really zero. So how can we improve upon the above model to explain this
reality? Deformed nucleon picture is an obvious solution. The calculations,
including deformations, are long and tedious [122]. However they are on the
right track. For smaller deformations PD (A) ∼ 0.04 for 28 Si and PD (A) ∼
0.02 for 48 Ca we obtain experimental values. This is highly satisfactory as it
demonstrates, indeed as remarked above, that the value of PD (A) decreases
with the A-number. Thus this is consistence with the expectation as to the
deformation of a nucleon within the EMC experimental results.
Quarks in Nuclei 457

12.6 Gamow-Teller (GT) Strength in Nuclei


As we saw in the previous chapter there is a sum rule relating the GT
strength for β− and β+ decays in nuclei within nucleonic degrees of freedom
as
2
Sβ − − Sβ + = 3gA (N − Z) (12.76)
48 90
As Sβ + strength is assumed to be zero for nuclei like Ca, Zr this
2
3 (N − Z) (in units of gA ) is a lower limit of the Sβ -strength. So far as a
generic property is concerned, one finds that the Sβ -strengths in nuclei are
quenched by 30 − 70%. These have been obtained experimentally in (p, n)
reactions.
The role of ∆-excitations in nuclei to explain this quenching of GT strength
has been emphasised [123]. Interestingly our model discussed above takes ac-
count of ∆-degrees of freedom in nuclei in a natural manner.
So what has our model here, which incorporates the quark structure of
nucleon in nuclei, has to say about the GT strength in nuclei? Using this
picture of a nucleus, as we saw in the previous section, |0 > is given by a
Slater determinant defined in Equation 12.37 and characterized by a strength
given by Equations 12.38 and 12.47. Now replacing the magnetic moment
operator by β− and β+ GT operators, we find:
A
X 3
X X
β± = ( σµ (i)τ+ (i)) (12.77)
a=1 a∈i=1 µ

Thus S N +∆ is obtained by using Equation 12.47. S ∆ , due to the reason


that there is no Pauli blocking effect by direct calculation, and subtraction of
the two gives S N , the strength residing at low energy in the nucleus. After
some lengthy algebra [124] one obtains the following GT strength for β− and
β+ cases:

SβN−+∆ = (6 + 3gA ) N + [3 + 3gA (1 − gA )] Z


 
∆ 2 1
Sβ − = gA rπN 4 N +Z
3
SβN− = SβN−+∆ − Sβ∆− (12.78)
and
SβN++∆ = (3 + 3gA ) N + [6 + 3gA (1 − gA )] Z
 
∆ 2 1
Sβ + = gA rπN ∆ N + Z
3
SβN+ = SβN++4 − Sβ∆+ (12.79)
458 Group Theory in Particle, Nuclear, and Hadron Physics

where
5
gA = (1 − PD (A)) (12.80)
3
and

   2
72 1 6 −1
rπN ∆ = 1 − PD (A) / 1 − PD (A) (1 + PD (A)) (12.81)
25 2 5

where rπN 4 is the so-called pion-nucleon-delta coupling constant and PD (A)


holds for a nucleon in a particular nucleus of mass number A.
The famous sum rule in Equation 12.76 has been the only tool to analyse
GT strengths discovered in the (p, n) reactions. One usually assumes Sβ + = 0
for nuclei. But above we saw that actually a considerable amount of strength
exists in the nucleus. Hence for 90 Zr, SβN+ = 52.1(in units of gA
2
) with PD (A) =
0.1. This is just an estimate to show that it is not justified to take Sβ + = 0.
This can be obtained empirically by (n, p) reactions planned at TRIUMF,
Canada.
So SβN− for 90 Zr with PD (A) = 0.1 we get 86.5gA 2
This is much larger
2
than the 3 (N − Z) = 30gA estimate. Hence a more systematic study of both
(p, n) and (n, p) reactions is required to clarify the muddled situation. What
is interesting above is that a deformed nucleon in a nucleus does give new
predictions for SβN− and SβN+ , where modifications of the properties of a nucleon
embedded in a nucleus – as per EMC experimental results – play a significant
role.

12.7 Solutions of Problems

Solution 12.1:
The Young Diagram for the 3-quarks each for both the groups SU (4)SF
and SU (3)C are,

⊗ ⊗ = ⊕ ⊕ ⊕
Now for 6-quarks we get products like below twice,
Quarks in Nuclei 459

1 1 1
1 1 1 1 1
⊗ 2 = 2 ⊕ 2 ⊕ 2

1
2 1 1
⊕ 1 ⊕ 1 2 ⊕ 1 2

And

⊗ = ⊕ · · · ( but no diagram )

And

⊗ = ⊕ · · · ( but no diagram )
However, the following product occurs twice

1 1 2
2 1 2 3 1 2
⊗ 3 = ⊕ 3 ⊕ 3

1
1 2 1
⊕ 2 3 ⊕ 3 ⊕ 2 3
So in taking the product of 3-quark terms, then counting all the terms in
(56)SU (4)SF and SU (3)c Young Diagrams for 6-quarks, these occur indepen-
dently five times each.
————————————————–
Solution 12.2:

To obtain the fully antisymmetric states 16 of SU (12)SF C reduced with
respect to the subgroup SU (4)SF ⊗ SU (3)c , we take product of two conjugate
Young Diagrams in the two subgroups to obtain the good fully antisymmetric
states. So it is
460 Group Theory in Particle, Nuclear, and Hadron Physics

= ( , )⊕( , )⊕( , )

⊕( , ) ⊕ ( , )


924 = (50, 1) ⊕ (64, 8) ⊕ (6, 27) ⊕ (10, 10) ⊕ 10, 10
————————————————–
Solution 12.3:
¯
 
One finds that 19 antisymmetric state in SU (12) has dimension 220
hence this antisymmetric state as per reduction in SU (4) ⊗ SU (3) space is


¯
(20)SU (4)SF ⊗ (1)SU (3)c

Let us take this tri-baryonic system as built up from the one-baryonic state
⊗ two-baryonic states. Then for the colour singlet component of the 9-quarks
colour singlet state, by using Equation 12.9 for the one-baryon states times
Equation 12.11 for the two-baryons states, the above is obtained from

( , ) ⊗ ( , )→( , )

¯ 1)
(20, 1) ⊗ (50, 1) → (20,
And the corresponding CFP is simply,
  21
[19 ][333][333 ˜ ] h[3] h[33]
C[13 ][3] 3̃ ,[16 ][33] 33 =
[] [˜] h[333]
Quarks in Nuclei 461

now

9!
[333] = : h[333] = = 42
3.4.5.2.3.4.1.2.3
q
5
So C = Multiplying √15 of the 1 baryonic term one obtains the colour
42 .
q
1
singlet part as 42 -
————————————————–
Solutionm 12.4:

3
X 3
< N| Q(i)2 |N >= [< φρN |Q(3)2 |φρN > + < φλN |Q(3)2 |φλN >]
i=1
2

1 1
< φρn |Q(3)2 |φρn >= (udd − dud)|Q(3)2 |(udd − dud) =
2 9
1 1
< φλn |Q(3)2 |φλn >= (udd + dud − 2ddu)|Q(3)2 |(udd + dud − 2ddu) =
6 3
4
< φρp |Q(3)2 |φρp >=
9
2
< φλp |Q(3)2 |φλp >=
9
So for neutron
3  
X 3 1 1 2
< n| Q(i)2 |N >= + =
i=1
2 9 3 3

And for proton


3  
X 3 2 4 2
< p| Q(i) |p >= + =1
i=1
2 9 9

And so
A
X X 2
< ψt | Q(i)2 |ψt >= Z + N
t=1 i
3
————————————————–
Chapter 13
Quark Gluon Plasma(QGP)

13.1 Basics for QGP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 463


13.2 Finite Lie Group Transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 467
13.3 Group Characters of the Lie Group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 471
13.4 Measure Function of SU(n) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 476
13.5 Symmetries and Partition Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 478
13.6 Colour Singlet and Coloured QGP States . . . . . . . . . . . . . . . . . . . . . . . 480
13.7 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 488

13.1 Basics for QGP


It is known that for sizes R ∼ 1f m, baryons and mesons confine quarks and
gluons within themselves. Indeed, unlike the bound state of a hydrogen nucleus
which may break up into its constituent electron and constituent proton above
13.6eV – quarks and gluons cannot be separated from their bound states. Thus
the nucleus is treated as composed of protons and neutrons and not of quarks.
Hence the deuteron is a loosely bound system of one proton and one neutron,
while He3 is composed of two protons and one neutron, and He4 is made
of 2 protons and 2 neutrons. So is the quark degree of freedom completely
irrelevant for the ground states of these nuclei?
The answer for most of the time is yes, but there are specific situations
where we still require quark degrees of freedom to comprehend nuclear phe-
nomena. In fact, we studied these in Chapter 12. Here we point out that the
charge distribution of nuclei is studied as a function of r, the relative distance
between nucleons, in the high energy Coulomb interactions between the in-
coming electrons and the target nuclei. Thus we can extrapolate to r → 0.
As individual nucleons have a finite size, this r → 0 density demands that for
the deuteron, (He3 , H 3 ) and He4 , the 6-quark, the 9-quark and the 12-quark
configurations, respectively, cannot be ignored. These were studied in terms of
the hidden colour components of multi-quarks state in the nuclei in Chapter
12.
Here we have pointed out that already quark gases of different kinds, 6,
9 and 12 quarks manifest themselves even in the ground state of nuclei in
which temperatures ∼ 0 and with small values of density. So in the heavy ion
collisions of nuclei, where high densities and high temperatures are attained,

463
464 Group Theory in Particle, Nuclear, and Hadron Physics

we may expect to see a quark gluon plasma. This simple expectation is borne
out by finer theoretical analysis [125] [126] [127].
However for individual nucleons within the ground states of nuclei, the
confinement size of quarks is ∼ 1f m. Furthermore, in high energy heavy ion
collisions when a putative quark gluon plasma (QGP) state has been formed,
the necessity of ensuring that no quarks or gluons leak out of the QGP implies
the expectation for the confinement size to reach ∼ 10 − 20f m. As Mueller
notes, “Thus the transition from hadronic matter to the QGP, is a transition
from local quark confinement to global quark confinement” [126].
We are using the word global for the above QGP size of ∼ 10 − 20f m.
Moreover, currently favoured scenarios envisage the existence of quark clusters
of sizes ∼ 1km within neutron-stars of dimensions ∼ 10km. Moreover, it is also
plausible that full-fledged quark-stars of sizes ∼ 10km may form at various
stages in the evolution of the universe. Clearly these large-sized entities should
likewise be globally colour singlet in nature [125].
Now local colour confinement, through understood phenomenologically,
has been extremely difficult to describe within QCD. It is safe to say that as
of now no conclusive proof of local colour confinement exists within QCD. Nor
is it clear which symmetry or which aspect of QCD permits the existence of
the large-scale correlations which are required for global colour confinement.
In this chapter we shall attempt to tackle this issue in some detail.
Note that the theory of strong interactions QCD is a locally exact gauge
theory. It has been very successful in explaining the strong interactions phe-
nomenon at high energies. At low energies phenomenological models, based on
some aspects of QCD, are doing reasonably well. But all these can be dubbed
as local-confinement in a hand waving manner. Yet the global-confinement is
clearly different from the local-confinement. Thus attempting to comprehend
global-colour-confinement using models based on local QCD may not be a
fruitful enterprise. Thus this shows that we should resort to the full global
SU (3)c group to understand this phenomenon. And indeed, this is what is
required, as we demonstrate here.
Hence at extremely high energy densities, when nucleons are dramati-
cally squeezed into individual sizes much smaller than ∼ 1f m, the other-
wise strongly bound and confined lower-level sub-hadronic particles may leak
out into the entire nuclear region. This utter obliteration of normally rigid
hadronic boundaries would, in essence, lead to the formation of an entirely
new form of matter consisting of the liberated quarks, anti-quarks and gluons.
This hypothetical novel state is generally denoted the Quark Gluon Plasma
(QGP). It is generally believed that this putative QGP must have existed for
a few micro-seconds after the Big Bang, at temperatures and densities far
exceeding those present in the universe today.
Empirically observing the theorized QGP is thus a veritable Holy Grail for
experimental physics, with a common approach involving attempts to partially
recreate the Promethean conditions of the early universe through ultra-high
Quark Gluon Plasma(QGP) 465

FIGURE 13.1: Schematic picture of high energy heavy ion collisons to create
QGP

energy heavy ion collisions. A schematic picture of how such a collision would
behave is provided in Figure 13.1.
Figure 13.1 is a schematic picture of a central heavy ion collision. As
strongly Lorentz-contracted lead and gold nuclei collide, these would prac-
tically pass through each other, thus leaving behind a residual low-baryon
number and high-energy density matter which we associate with QGP.
Various signature signals have been suggested to prove the existence of
QGP in the laboratory. A few of these are : (1) strangeness enhancement;
(2) photon and dilepton spectra; (3) J/ψ melting; and (4) hybrid formation
[128],[129]. However hydrodynamic properties such as elliptic flow and jet
quenching have proven to be far more discriminating experimental signatures
of the QGP.
Work to detect such a putative QGP began in earnest in the 1980s and
1990s. The two main centres at the time were the Relativistic Heavy Ion Col-
lider (RHIC) at the Brookhaven National Laboratory in USA and the Super
Proton Synchrotron (SPS) at CERN in Geneva, Switzerland. At Brookhaven,
silicon and gold ions were already being accelerated to 10 GeV per nucleon,
while CERN commenced with 160 GeV per nucleon beams of sulfur and lead.
In the former locality, with high-energy beams colliding into stationary
targets, only modest centre-of-mass energies of the order of 5−17 GeV/nucleon
were obained. The relevant energy to extract information about matter is the
center-of-mass energy. The centre-of-mass energy for fixed targets grows only
as the square root of the beam energy, thus giving an advantage to RHIC in
466 Group Theory in Particle, Nuclear, and Hadron Physics

the long run. However, after working for almost 15 years the CERN scientists
announced indirect evidence of a “new state of matter” (commonly identified
as QGP) in 2000. This was confirmed in 2005 by experiments performed on
(gold + gold) at RHIC.
Note that for the entire 30 years or so the experimental programmes at
CERN and BNL had been motivated by the exciting expectation of a gaseous
QGP (i.e., a gas of quarks, anti-quarks and gluons) obtainable at high den-
sities and high temperatures ∼ 170M eV . Though the experimental dream
ultimately fructified in the years from 2000 to 2005, the final results sent
shock waves throughout the field of Theoretical Physics, demolishing many
a once-respected model of QGP. Further empirical work in this line contin-
ues to yield new results, which continue to act as cataclysmic after-shocks in
the domain of theoretical models, shaking up entire frameworks and forcing
the development of new models. Below we summarize the highly fluid and
dynamic situation.
In these experiments, there are clear indications of the role of quarks in
determining the bound state distributions of the QGP produced. However,
as a Zajc has noted, “. . . there is also compelling evidence that the matter
does not behave as a quasi-ideal state of free quarks and gluons. Rather, its
behaviour is that of a dense fluid with very low kinetic viscosity exhibiting
strong hydrodynamic flow” [130].
Thus the RHIC data posed a very basic and highly disturbing theoretical
problem. As Heinz aptly observed, “It is impossible to understand the collec-
tive flow data by assuming a weakly interacting gaseous plasma. The QGP is
a strongly coupled liquid. In fact, it is the most perfect liquid ever created in
the laboratory” [131].
We find that the hydrodynamic model in describing the hadron production
data in Au + Au collisions at the RHIC has dramatically forced upon us a
paradigmatic change in our understanding of QGP. And it is that “. . . instead
of behaving like a gas of weakly interacting quarks and gluons, as naively
expected on the basis of asymptotic freedom in QCD, its collective perspective
rather reflect those of a ‘perfect fluid’ with (almost) vanishing viscosity” [132].
In QGP technical jargon, the strong elliptical flow as defined by a parame-
ter v2 is regarded as a function of the transverse momentum pT of the emitted
particles. Over the relevant physical region
p where v2 is plotted as a function of
the transverse kinetic energy K.E. = m2 + p2 − m, one observes a surpris-
ing scaling behaviour in QGP. This v2 (K.E T ) demonstrates that the observed
baryons and the observed mesons follow two different branches. Upon further
scaling of both K.E.T and v2 (K.E.T ) by nq , the above two different branches
merge into a universal scaling curve for elliptic flow [133]. φ meson although
more massive than a nucleon, follows the scaling curve of mesons and not that
of baryons. Here it is imperative to quote an esteemed authority, “While it is
tempting to conclude that the strict scaling according to constituent quarks
content provides incontrovertible evidence for the underlying role of quark
Quark Gluon Plasma(QGP) 467

degrees of freedom in establishing the elliptic flow, such a conclusion appears


to be at odds with the observation of perfect fluidity” [130].
Lifetimes of quasi-particles are of the same order as the long time over
which hydrodynamic flow persists. This is utterly incompatible with the very
short mean free paths implied by low viscosity and is hard to reconcile with the
strong empirical evidence for quark number scaling of the flow phenomenon.
So the great conundrum of QGP is this – is it a fluid or a gas? If both,
then how can one explain this fundamental Manichaean duality? A reasonable
resolution of this puzzle does not exist at present. However, we shall demon-
strate how symmetry principles play a fundamental role in explaining this
perplexing problem in QGP.

13.2 Finite Lie Group Transformations


There is an issue in the application of the continuous symmetries in par-
ticle, nuclear and hadron physics. And that is whether it is sufficient to con-
fine oneself to the Lie algebra only [118]. Many scientists think this is good
enough. The admirable success of gauge theories in the Standard Model of
strong, weak and electromagnetic interactions built around the Lie algebra of
SU (3)c × SU (2)L × U (1)Y has cemented this view.
Hence, for example, the issue of the confinement of quarks in hadrons is
being vigourously pursued within the field of QCD, the exact gauge theory
of the strong interaction. But as we saw, the issue of QGP is actually how
one goes from the local colour confinement to the global colour confinement.
Over a range of ∼ 1 f m the local colour confinement is clearly an issue
for QCD to handle. But what is this global colour confinement, say over a
range of ∼ 10 − 20f m ? This is global, as it exists over a larger region. Global
colour confinement is also believed to be relevant for quark-stars of dimensions
∼ 10 km and of ∼ 1km inside a neutron star. In order to avoid the splitting
of free quarks and gluons globally, it is common to demand that global colour
singletness be valid over these large regions [126],[125].
So what holds together these large colour-singlet entities? Evidently, Lie
algebra, being purely local in nature, may not be of much applicability here.
Thus the global colour-singletness in QGP and quark-stars demands the use of
the complete global Lie group, of which QCD is the local algebraic counterpart.
Clearly there exists a duality in the phenomenon of confinement of quarks and
gluons, one particular and regionalised local colour confinement within the
∼ 1 f m scale, and another distinct and far-reaching global colour confinement
from ∼ 10 f m to ∼ 10 km or so.
For the SU (3)c QCD group we applied Lie algebra with the notation of
[λc , λj ] = 2ifijk λk with λi generators defined in the standard way. However to
proceed to the full SU (3)c Lie group these do not provide the most convenient
468 Group Theory in Particle, Nuclear, and Hadron Physics

basis. For that purpose we need to redefine new generators as below. This
requires complicated notation. Thus for the sake of simplicity we follow the
conventions of Greiner and Mueller [32].
For the unitary group U (n), the n2 generators Cim (i, m = 1, 2, . . . n) satisfy
the following commutation relation:

[Cim , Cjm ] = δjm Cin − δin Cim (13.1)


where the matrix

(Cim ) = (δim ) (13.2)


i.e., with 1 at the location of ith row and mth column and zero elsewhere. Now
an arbitrary element of group U (n) is:
X
exp(−i θk,l Ck,l ) (13.3)
k,l
2
where θk,l is the angle in the n group parameters.
To restrict ourselves to the SU (n) group we construct traceless matrices.
Cim for i 6= m are automatically traceless by definition. For i = m cases as
the unit matrix 1 commutes with all Cim , we define these as traceless. Note
that a tilde on a generator indicates its tracelessness. Hence,
1
C˜ii = Cii − (13.4)
n
 
−1/n
 . −1/n
C˜ii = 

 (13.5)
 1 − 1/n 
−1/n
Note that (1 − 1/n) is the (ith , ith ) element above.
Clearly it is traceless and by its definition ΣC˜ii = 0. This condition implies
that only (n-1) of these function are linearly independent and so there are
(n2 − 1) generators for SU (n).
Now for SU (3) we have two sets of generators: Cim = δim and the standard
λi matrices. These are related as follows:
 
0 1 0
λ1 = 1 0 0 = C21 + C12
0 0 0
 
0 −i 0
λ2 =  i 0 0 = i(C21 − C12 )
0 0 0

(check C21 + C12 + i · i(C21 − C12 ) = 2C12 = λ1 + iλ2 )


Quark Gluon Plasma(QGP) 469

 
1 0 0
λ3 = 0 −1 0 = C11 − C22
0 0 0
 
0 0 1
λ4 = 0 0 0 = C31 + C13
1 0 0
 
0 0 −i
λ5 = 0 0 0  = i(C31 − C13 )
i 0 0
 
0 0 0
λ6 = 0 0 1 = C32 + C23
0 1 0
 
0 0 0
λ7 = 0 0 −i = i(C32 − C23 )
0 i 0
 
√ 1 0 0 √
λ8 = 1/ 3 0 1 0  = 1/ 3(C11 + C21 − 2C33 ) (13.6)
0 0 −2
In U(3),

3
X 3
X X
exp(−i θkl Ckl ) = exp(−i θkl Ckl − i θkk Ckk ) (13.7)
kl=1 k6=l=1

To transform to a traceless form, note Ckl = C˜kl (k 6= l) and from Equation


13.4 Cii = C˜ii + n1 1. Hence,
3 3 3
˜ )exp(− i 1
X X X
= exp(−i θkl C˜kl − i θkk Ckk θkk ) (13.8)
n
k6=l=1 k=1 k=1

The right-hand term is simply the U (1) phase factor and using U (3) =
SU (3) × U (1), the rest of the terms correspond to the group SU (3). Using
Equation 13.8 we restrict from U (3) to SU (3) by demanding Σ3k=1 θkk = 0.
Note that although we have used Cim form of generators here, in the
original SU (3) Lie algebra we had used λi (i = 1, ..8) generators and the quarks
were required to reflect eigenstates of the corresponding Cartan subalgebra i.e,
with the diagonal generators λ3 and λ8 . As it is the same quarks that we shall
talk of here, we need to connect the weights of SU (3) algebra in terms of
λi and Cim . Given Σ3k=1 θkl = 0, we use the C˜ii term from Equation 13.4 in
Equation 13.7. Now let us look at diagonal generators to start with,
470 Group Theory in Particle, Nuclear, and Hadron Physics

φ ψ
exp(−iΣk θlk (Ckl − 1/n)1) = exp(−i( λ3 + √ λ8 )) (13.9)
2 3
with

1 = C11 + C22 + C33 (13.10)


we get

θ11 C11 + θ22 C22 + θ33 C33 − (1/3)(θ11 + θ22 + θ33 )(C11 + C22 + C33 )

φ ψ
= λ3 + √ λ8 (13.11)
2 3
with the additional condition θ11 + θ22 + θ33 = 0(mod(2 π)
i.e., = 0, 2π only as θ11 , θ22 , θ33 → (−π → π),

1 1
[θ11 − (θ11 + θ22 + θ33 )]C11 + [θ22 − (θ11 + θ22 + θ33 )]C22 + [θ33
3 3
1
− (θ11 + θ22 + θ33 )]C33
3
φ ψ 1
= (C1 − C2 ) + √ √ (C1 + C2 − 2C33 ) (13.12)
2 3 3
So,
1 φ ψ
θ11 − (θ11 + θ22 + θ33 ) = + (13.13)
3 2 3
1 φ ψ
θ22 − (θ11 + θ22 + θ33 ) = − + (13.14)
3 2 3
1 2
θ33 − (θ11 + θ22 + θ33 ) = − ψ (13.15)
3 3
Hence,

φ φ
θ11 − θ22 = φ, θ11 − θ33 = + ψ, θ22 − θ33 = − + ψ (13.16)
2 2
For the infinitessimal transformation of the diagonal elements we have,

 
1 − iθ11 0 .. 0
X  0 1 − iθ22 .. 0 
1−i θkk Ckk =
 ..
 (13.17)
.. .. .. 
k
0 0 .. 1 − iθnn
Quark Gluon Plasma(QGP) 471

Now, exponentially for finite transformations,


 −iθ 
e 11
X  e−iθ22 .. 
exp(−i θkk Ckk ) =   (13.18)
 .. 
k −iθnn
e
Let us use k = exp(−iθkk ) in this matrix for convenience.
The above was for the diagonal matrix. For the definition in Equation 13.6
for the non-diagonal λ matrices
λij = Cij + Cji , λji = i(Cji − Cij ), i < j (13.19)
so
λ1 → λ12 = C12 + C21 , λ2 → λ21 = i(C21 − C12 ) (13.20)
For infinitesimal transformations one therefore has:
X X X X
1−i θkk Ckl = 1 − i θkk Ckl − i θkl λkl − i θlk λlk
kl k k<l k<l

1 − iθ˜11 θ˜12 − iθ˜21


 
..
θ˜12 + iθ˜21 1 − iθ˜22 
=  (13.21)
 . .. 
...
and

θ̃kl + iθ̃lk = θkl , θ̃kl − iθ̃lk = θlk , k < l and θ̃kl = θkl (13.22)

13.3 Group Characters of the Lie Group


Let us consider a matrix D(ga ) representing a group element ga < G. Then
this is not unique as a similarity transformation SD(ga )S −1 , S ⊂ D(g) yields
an equivalent form of the same representation. So how do we describe the
invariant properties of a group? One method is to choose eigenvalues of the
representation which do not change under a similarity transformation. So far
we have used the Casimir operator to classify the representations and this is
consistent as the Casimir operator commutes with all the generators of a Lie
group.
Another such operator is the character of a representation, which is de-
fined as the sum of the diagonal generators:
n
X
χ(ga ) = Dii (ga ) (13.23)
i=1
472 Group Theory in Particle, Nuclear, and Hadron Physics

where n is the dimension of the matrix representation. This is invariant under


a similarity transformation,

n
X X X
0 0
χ (ga ) = Dii (ga ) = Sij Djk (ga )(S −1 )ki = Djk (S −1 S)kj
i=1 i,j,k jk

X
= Djj (ga ) = χ(ga ) (13.24)
j

These have been extensively used to study irreducible representations of


finite groups [41]. But our interest in particle, nuclear and hadron physics is
in the continuous Lie groups, and that is what we define a character of here.
First characters of the group SO(3) in an irreducible representation can be
derived quite simply. We know that the spherical harmonic function Yml (θφ)
provides the (2l+1) dimensional basis of the full rotation group. It is an ir-
reducible basis. The character of this group will specify different irreducible
representations. Thus the character will differ for different values of l and thus
distinguish the different irreducible representations of the group SO(3).
Next note that if we take the group element,
 
cos φ − sin φ 0
R3 (φ) =  sin φ cos φ 0 (13.25)
0 0 1
Let us take the similarity transformation RR3 (φ)R−1 = Rk (φ) which shifts
the axis to an arbitrary direction and specifies a rotation by the same angle.
Hence the angle φ specifies a class of rotation. This implies that, given an
angle φ to specify a rotation along any axis in SO(3), then a similarity trans-
formation to another axis does not change its angle φ and hence this shall
define a class.
Now define a rotation by α around the 3-axis and let it act on the spherical
harmonic function,

R3 (α)Yml (θ, φ) = Yml (θ, φ − α) = e−imα Yml (θ, φ) (13.26)


Thus the representation of such a rotation is the diagonal matrix,
 −ilα 
e
e−i(l−1)α
Dj (R3 (α)) = 
 
 (13.27)
 .. 
eilα
Whence the character is
l
X
χl (α) = T rDj (R3 (α)) = e−imα (13.28)
m=−l
Quark Gluon Plasma(QGP) 473

2l
−ilα
X ei(2l+1)α − 1 sin(2l + 1) α2
l
χ (α) = e eikα = e−ilα =

e −1 sin α2
k=0

The character here holds for the class of odd-l states. The same holds for
even angular momentum cases as well. Here we may use ’j; instead of l above,
as a more general expression:

sin(2j + 1) α2
χj (α) = (13.29)
sin α2
Thus for j = 1/2 we have χ1/2 (α) = 2 cos α/2 and for j = 1 we have
χ1 (α) = 2 cos α + 1

————————————————–
Problem 13.1:
Verify j=1/2 and j=1 characters of the rotation group SO(3) through the
dj matrices of the rotation group (D-function).
————————————————–

Now we develop a general formalism for group characters of unitary groups.


Let a group element be given as:
X
gα = exp(−i θij Cij ) (13.30)
i,j

Then the character is:

χ(gα ) = T r(gα ) (13.31)


As the trace is invariant under similarity transformations Sgα S −1 , so every
unitary matrix can be transformed into diagonal form. Now we show that the
generator in the Cartan subalgebra specifies all representation matrices in the
diagonal form as
n
X
exp(− θkk Ckk ) (13.32)
k=1

It is sufficient to consider group elements in this diagonal form. Let the


weights or (eigenvalues) of Ckk (k = 1, 2, ...n) be given by (w1 , w2 , ...wn ), then
we can write the character of the representation as the sum over all possible
weights in:
X X
χ(α) = exp(−i θkk wk ) (13.33)
w1 ,..wn k

Note that θ11 , θ22 , ....θnn yield all the classes of a particular representation
α. One can solve for these characters in a most general form for U (n) and
474 Group Theory in Particle, Nuclear, and Hadron Physics

SU (n). For U(n) the character is specified by ( h1n , . . . ,P


hnn ) when maximal
weights w1 = h1n , w2 = h2n , · · · wn = hnn . For SU(n) k θkk = 0 and thus
the representation of SU(n) is characterized by:

h1n − h2n , h2n − h3n , · · · h(n−1)n − hnn , 0 (13.34)


We refer the reader to [32] for details. Here we just obtain the results of
χ(α) for SU (3) which is of primary interest to us. In SU (3) representations
are given as:

(p, q, 0) = (h13 − h21 , h23 − h31 , 0) (13.35)


Here h13 is the number of boxes in the first row of the Young Diagram of
the representation, h23 is the number of boxes in the second row of the Young
Diagram of the representation, and h33 is the number of boxes in the third
row of the Young Diagram of the representation.
In SU (3) we saw θ11 − θ22 = φ, θ11 − θ33 = φ/2 + ψ, θ22 − θ33 = −φ/2 + ψ,
with θ11 + θ22 + θ33 = 0
The character for SU (3) is:

(h −h23 ,h23 −h13 ,0) 2iψ


χSU13(3) = exp[ (h13 + h23 + h33 )]
3

h13 h23
X X h12 0 + 1 − h22 0 φ −1
= exp[−iψ(h012 + h22 0 )] · sin[( )φ] · sin( )
2 2
h12 0=h23 h23 0=h33
(13.36)

Example 1: Character for the singlet state, the Young Diagram is

(0, 0, 0) = (h13 − h23 , h23 − h33 , 0), so h13 = 1, h23 = 1, h33 = 1.


Putting in Equation 13.36

(0,0,0) φ φ
χSU (3) = exp(2iψ)exp(−2iψ) sin( ).(sin( )−1
2 2
=1 (13.37)

Example 2: Character of the fundamental representation, i.e., the triplet,


the Young Diagram is
Quark Gluon Plasma(QGP) 475

(1, 0, 0) = (h13 − h23 , h23 − h33 , 0), so h13 = 1, h23 = 0, h33 = 0. And

1
(1,0,0) 2iψ X h0 + 1 − 0 φ
χSU (3) = exp( ) exp(−iψ(h012 + 0)) · sin[( 12 )φ](sin( ))−1
3 0
2 2
h12 =0

= e2iψ/3 (1 + e−iψ sin φ(sin(φ/2))−1 )


= e2iψ/3 + 2e−iψ/3 (cos(φ/2)) (13.38)
or

= e2iψ/3 (1 + e−iψ (eiφ/2 + e−iφ/2 ))

= e(i(θ11 +θ22 −2θ33 )/3) (1 + e−i(θ22 −θ33 ) + e−i(θ11 −θ33 ) ) (13.39)


So

(1,0,0)
χSU (3) = ei(θ11 +θ22 +θ33 −2θ33 )/3 (1 + e−i(θ22 −θ33 ) + e−i(θ11 −θ33 ) )

= e−iθ11 + e−iθ22 + e−iθ33 (13.40)

(1,0,0)
X
χSU (3) = e−iθii (13.41)
i=1

————————————————–
Problem 13.2: Determine the character for the anti-triplet (0,1,0) repre-
sentation of SU (3).
—————————————————

Example 3 Character of the adjoint representation given by the Young


Diagram

(1, 1, 0) = (h13 − h23 , h23 − h33 , 0), so h13 = 2, h23 = 1, h33 = 0


(note we use θ1 = θ11 etc.).

2 1
(1,1,0)
X X 0 0
χSU (3) = e2iψ ( e−iψ(h12 +h22 ) · (sin((h012 + 1 − h022 )/2)φ)(sin(φ/2))−1 )
h012 =1 h022 =1

2
X 0 −1
= e2iψ ( e−iψh12 · (sin((h012 + 1)/2))(sin(φ/2))
h012 =1
476 Group Theory in Particle, Nuclear, and Hadron Physics
0
+e−iψ(h12 +1) (sin((h012 /2)φ)(sin(φ/2))−1 )
= e2iψ (e−iψ 2 cos(φ/2)+e−2iψ +e−2iψ sin(3φ/2)(sin(φ/2))−1 +e−3iψ 2 cos(φ/2))
= 2 + 2 cos(φ/2) + eiψ (ei(φ/2) + e−i(φ/2) ) + e−iψ (ei(φ/2) + e−i(φ/2) )
= 2 + eiφ + e−φ + ei(ψ+(φ/2)) + ei(ψ−(φ/2)) + e−i(ψ−(φ/2)) + e−i(ψ+(φ/2))
= 2 + ei(θ1 −θ3 ) + ei(θ2 −θ3 ) + e−i(θ1 −θ3 ) + e−i(θ2 −θ3 ) + ei(θ1 −θ2 ) + e−i(θ1 −θ3 )
= 2 + Σ3i<j ei(θi −θj ) + Σ3i<j e−i(θi −θj )
3
(1,1,0)
X
χSU (3) = 2 + 2 cos(θi − θj ) (13.42)
i<j

= 2 + 2(cos φ + cos((φ/2) + ψ) + cos(−(φ/2) + ψ)) (13.43)

————————————————–
Problem 13.3: Determine the character of a 27-plot representation.
————————————————–
Unsolved Problem 13.1: Demonstrate that the character of the 64-plet
representation (3,3,0) is given as:
(3,3,0)
χSU (3) = 4 cos φ(cos(3ψ + φ/3) + cos(3ψ − φ/2) cos(ψ + φ/2) + cos(ψ − φ/2))
+(1 + 2 cos φ)((2 + 2 cos φ) cos 2ψ + 1 + 2 cos(3φ/2)(2 cos ψ + cos φ/2))
+2[cos(φ/2 + ψ) + cos(−φ/2 + ψ) + cos(3φ/2)(cos(φ/2 + 2ψ)
+ cos(−φ/2 + 2ψ) + cos φ/2) + (1 + cos 3φ)]

————————————————–

13.4 Measure Function of SU(n)


In evaluating the characters of different irreducible representations of SU(n), we
obtain these as functions of the angle φ in SU (2) and of φ and ψ in SU (3). There may
be other relevant functions of the group SU(n). Let us denote these as f (SU (n)).
We would therefore need to be able to integrate these functions. In generalising say
the integration over the 3-dimensional (r, θ, φ) space, we define the corresponding
integration over the group SU(n) as:
Z
1
f (SU (n))µ(SU (n))dθ (13.44)
V
where dθ = dθ11 , dθ22 , · · · , dθnn are the group parameter. Variable µ(SU (n)) is
defined as the measure function of group SU (n). The integration is weighted by the
“volume” term,
Z
V = µ(SU (n))dθ (13.45)
Quark Gluon Plasma(QGP) 477

the orthogonality solution of characters of finite group carries over to the continuous
Lie groups. As an example of f (SU (n)) in above, we have in terms of the characters
of SU (n).
Z
1 ∗
χα χβ µ(SU (n))dθ = δαβ (13.46)
V
or
Z
1 0 0 ∗
χ(p+q) χ(p +q ) µ(φ, ψ)dφdψ = δpp0 δqq0 (13.47)
V
For the general expression of the measure function of SU (n) we refer the readers
to [32] and hence define it as,

µ(U (n)) = Πn
i<j |i − j |
2
(13.48)
where

k = e−iθkk (13.49)
and for SU (n) impose the condition θ11 + θ22 + ...θnn = 0.

Example 1 For U(2)


2
µ(U (2)) = |1 − 2 |2 = |e−iθ11 − e−iθ22 |

2
= |e−iθ11 /2−iθ11 /2−iθ22 /2+iθ22 /2 − e−iθ22 /2−iθ22 /2+iθ11 /2+iθ11 /2 |

2 2
= |e−(i/2)(θ11 +θ22 ) | |e−(i/2)(θ11 −θ22 ) − e(i/2)(θ11 −θ22 ) |

2 θ11 − θ22
= |e−(i/2)(θ11 +θ22 ) | 4 sin2
2
For SU(2) , θ11 + θ22 = 0.

θ11 − θ22 φ
µ(SU (2)) = 4 sin2 = 4 sin2 (13.50)
2 2
Thus Z 4π
V = µ(SU (2))dφ = 8π (13.51)
0
and
Z Z 4π
1 1
µ(φ)f (φ)dφ = (sin2 (φ/2))f (φ)dφ (13.52)
V 2π 0
(−π ≤ θ11 ≤ π and so − 2π ≤ (θ11 − θ22 ) ≤ 2π as θ11 + θ22 = 0)

Example 2 For U (3)

µ(U (3)) = |1 − 2 |2 |1 − 3 |2 |2 − 3 |2


2 θ11 − θ22 2 θ11 − θ33
= 4|e−i(θ11 +θ22 ) | sin2 ( ) · 4|e−i(θ11 +θ33 ) | sin2 ( )
2 2
478 Group Theory in Particle, Nuclear, and Hadron Physics
2 θ22 − θ33
·4|e−i(θ22 +θ33 ) | sin2 ( )
2
whence

θ11 − θ22 θ11 − θ33 θ22 − θ33


µ(SU (3)) = 64 sin2 ( ) sin2 ( ) sin2 ( ) (13.53)
2 2 2

= 64 sin2 (φ/2) sin2 (1/2((φ/2) + ψ)) sin2 (1/2(−(φ/2) + ψ)) (13.54)


where
Z π Z π
V = d(φ/2)d(ψ/3)µ(SU (3)) (13.55)
−π −π

(−π ≤ θ33 ≤ π; θ11 + θ22 − 2θ33 = 2ψ; so − 3π ≤ ψ ≤ 3π as θ11 + θ22 + θ33 = 0)

V = 24π 2 (13.56)
Thus we get
Z
1
µ(SU (3))f (φ, ψ)d(SU (3))
V
Z π Z π
8 φ ψ φ 1 φ
= d( )d( ) sin2 ( ) sin2 ( ( + ψ))
3π 2 −π −π 2 3 2 2 2
1 φ
sin2 ( (− + ψ))f (θ, φ) (13.57)
2 2

————————————————–
Problem 13.4: For SU(3) demostrate that V = 24π 2 .
————————————————–

13.5 Symmetries and Partition Functions


Given a thermodynamics ensemble in thermal equilibrium, all relevant physical
properties can be obtained from the partition function defined as:

Z = T r(e−BH ) = Σi e−βEi (13.58)


Hence the pressure P and internal energy are [32],

∂ ∂
P =T (lnZ), E = T 2 (lnZ) (13.59)
∂V ∂T
Assuming a non-interacting gas of massless fermions and bosons, one analytically
obtains the following results,

7π 2 π2
lnZF = gF ( )V T 3 , lnZB = gB ( )V T 3 (13.60)
120 30
Quark Gluon Plasma(QGP) 479

where the independent degrees of freedom for fermions, gF = 2 · 3 · 2 = 12 (for


2-flavours) and for bosons, gB = 2 · 8 = 16.
Thus we can study QGP using these equations. However these results do not
form any constraints on colour space, i.e., these hold for the case without any colour
degree of freedom. However as QGP is expected to be globally colour singlet, the
above results have to be taken with a pinch of salt. Therefore we are really interested
in those partition functions which are constructed to be globally colour singlet only.

Zsinglet = T rsinglet (e−BH ) (13.61)


Below we show how to do it in a consistent manner for the SU (3)c colour group.
First we realize that H in Equation 13.61 conserves SU (3)c symmetry. It does
not have enough colour freedom to distinguish between different representations of
SU (3)c . Hence we add a sum over the weight operator Cˆkk of the SU (3) as the trace
in Equation 13.58,
X
Z = T r(exp(−βH − i θkk Ckk )) (13.62)
k

The trace is over the diagonal of all possible states. In SU (3)c we la-
bel different irreducible representations by their dimensions α = (p, q) such as
(0, 0), (0, 1), (1, 0), (1, 1) etc.. Different members of each of these representations are
distinguished by β = (tc3 , y c ), i.e., the eigenvalue of the two diagonal generators as
per the Cartan subalgebra in the colour space SU (3)c . Thus,
X X
Z= < α, β|exp(−βH − i θkk Ckk )|α, β > (13.63)
α,β k
XX 0 0 0 0
= < α, β|exp(−βH)|α , β >< α , β |exp(−iΣk θkk Ckk )|α, β > (13.64)
α,α0 β,β 0

Next note that H being independent of colour is diagonal in β. Also Ckk being
diagonal generators of the SU (3)c group do not change the irreducible representation
α and thus leave their different states β unaltered as well:

X X X
Z= < α|e−BH |α > < α, β|exp(−i θkk Ckk )|α, β > (13.65)
α β k

Note that the second term is the character of the representation α = (p, q),

(p,q)
X X
χα
α (θkk ) = χ(p,q) (θ, ψ) = < α, β|exp(−i θkk Ckk )|α, β > (13.66)
β k

Now,
1 X
< α|e−BH |α >= < α, β 0 |e−BH |α, β 0 > (13.67)
d(p, q) 0
β

where d(p, q) is the dimension of the irreducible representation α, e.g., d(1, 1) = 8 and
so forth. We have weighted the matrix element of e−βH above as it is independent
of the β = (tc3 , y c ) quantum numbers of SU (3)c .
In the above expression the trace is over β 0 , which is all the members of the
480 Group Theory in Particle, Nuclear, and Hadron Physics

irreducible representation of α = (p, q). This is the partition of a particular repre-


sentation that we have been seeking [134] [135]:

Zα = Z(p,q) = T r(p,q) (exp(−βH)) (13.68)


Thus,
X Zp,q
Z= χ(p,q) (φ, ψ) (13.69)
p,q
d(p, q)
Using the orthogonality property of the characters and substituting Equation
13.66 in Equation 13.69 we get,
Z
Z(p,q) = d(p, q) dφdψµ(θ, ψ)χ∗( p, q)(θ, φ)Z(p, q, T, V ) (13.70)

where µ(θ, ψ) is the measure function of the group derived from the orthogonality
of the characters. We have labelled Z(p, q, T, V ) indicating that it shall be found to
depend upon T and V of the confined QGP that it shall represent.
Z(p, q) in Equation 13.73 is what we have to evaluate using the corresponding
characters and Z(p, q, T, V ) functions. For Z(0,0) , i.e., colour singlet, χ(0,0) = 1 and
the function simplifies considerably.

13.6 Colour Singlet and Coloured QGP States


Now we have Equation 13.62 as:

Z(θ, φ, T, V ) = T r[exp(−βH + iφtc3 + iψy c )] (13.71)


where tc3 c
and y are the diagonal generators of the Abelian subgroup, as Cartan
subalgebra of SU (3)c . Our plasma consists of light spin-1/2 quarks and antiquarks
in triplet (1, 0) and anti-triplet (0, 1) colour representations and massless spin-one
gluons in octet representations (1, 1).
The non-interacting Hamiltonian H is diagonal in the occupation number repre-
sentation. In the same representation we can express the charge operators tc3 and y c
as linear combinations of particle number operators. Thus we can calculate Z in the
occupation number representation. With an “imaginary chemical potential” this is
just like a grand canonical partition function for free fermions and bosons. Hence
one obtains for quarks:

Zquark = Π∞
k=1 Πq=r,b,g [1 + exp(−βk − iαq )][1 + exp(−βk + iαq )] (13.72)

where the first term is for particles (0, 1) and the second is for anti-particles. k are
the single particle values. And for gluons,

Zgluons = Π∞
k=1 Πg=µ,ν,ρ,σ [1 − exp(−βk + iαg )]
−1
[1 − exp(−βk − iαg )]−1 (13.73)

Let us calculate αq and αg from the weight diagrams, given in Figure 13.2.
Quark Gluon Plasma(QGP) 481

yc yC
2
+
3
+
^3
C C
1q t
3

1 +
1
1
+
i
2 2 ^
2 2
1
2^ 3
3

(a) (b)

yc

+1

c
t
-1 +1
1 O
!_
+
2 2

-1

(c)

FIGURE 13.2: Weight diagram of the triplet, antitriplet, and octet repre-
sentations used to calculate partition functions
482 Group Theory in Particle, Nuclear, and Hadron Physics

Therefore for the triplet (1, 0),


2
αr = φ/2 + ψ/3, αg = −φ/2 + ψ/3, αb = − ψ (13.74)
3
and antitriplet,
2
αr̄ = −φ/2 + ψ/3, αḡ = φ/2 − ψ/3, αb̄ = ψ (13.75)
3
and for the octet (1, 1),

αµ = αr − αg = φ, αν = αg − αb = −φ/2 + ψ,
αρ = αr − αb = φ/2 + ψ, ασ = 0 (13.76)
and its complex conjugate yield the other four quantum numbers.
In the occupation number representation, the generating function for quarks and
gluons can also be written as [135]

X X
Zquark = exp( log[1 + exp(−βj + iαq )]) (13.77)
j=1 q=r,g,b

X X
Zgluons = exp(− log[1 − exp(−βj + iαg )]) (13.78)
j=1 g=µ,νρ,σ

Now the total generators function is

Z(φ, ψ, T, V ) = Zquark (φ, ψ, T, V )Zgluons (φ, ψ, T, V ) = eX (13.79)


where


X X ∞
X X
X= log[1 + exp(−βj + iαq )] + log[1 + exp(−βj − iαq )]
j=1 q=r,g,b j=1 q=r,g,b


X X ∞
X X
− log[1− exp(−βj +iαg )]− log[1− exp(−βj −iαg )] (13.80)
j=1 g=µ,ν,ρ,σ j=1 g=µ,ν,ρ,σ

α X
α
X e−βnj X e−βnj X
X= ((−1)n+1 2 cos(nαq )+ 2 cos(nαg )) (13.81)
j=1 n=1
n n g=µ,ν,ρ,σ
q=r,g,b

Continuous approximation for single particle energy level yields,


Z ∞ Z ∞
X V 2
→ ρ()d() = d (13.82)
g 0 0 2π 2
Hence,
∞ ∞
V 2 −βnj V T3
X Z
e−βnj = dq g e d = dq g (13.83)
j=1 0 2π 2 π5
Quark Gluon Plasma(QGP) 483

where dq = (2S + 1)(2I + 1) = 2 × 2 = 4 for quarks and dg = 2 for gluons (where 2


is from polarization).
Putting Equation 13.83 in Equation 13.81, we obtain,


X 2V T 3 X 2V T 3 X
X= ((−1)n+1 dq cos(nα q ) + dg cos(nαg )) (13.84)
n=1
π 2 n4 π 2 n4 g=µ,ν,ρ,σ
q=r,g,b

π2 V T 3 π2 V T 3 X αq αq
X= (dq(7/10) + dg(8 · 4)/30) + (dq (1/2)( )4 − ( )2 )
12 12 q
π π

2
X |αg | 1 |αg | 4
+dg (−1/2 + ( − 1) − ( ) ) (13.85)
g=µ,ν,ρ,σ
π 2 π

Z(φ, ψ, T, V ) = eX = Z (0) (T, V )Z (1) (φ, ψ, T, V ) (13.86)


where

π2 V T 3 7 16
Z (0) (T, V ) = exp( ( dq + dg)) (13.87)
12 10 15
This is the partition function for an ideal quantum gas in the Stefan-Boltzmann
limit.

π2 V T 3 X 1 αq 4 αq
Z (1) (φ, ψ, T, V ) = exp( (dq ( ( ) − ( )2 )
12 2 π π
q=r,g,b
2 4
X |αg | |αg |
+dg (−1/2 + ( − 1) − 1/2( − 1) )) (13.88)
g=µ,ν,ρ,σ
π π
These two form the essential components for performing the integration to de-
termine different irreducible representation partition functions, including the colour
singlet component as well. Putting these in Equation 13.70,
Z
Z(p,q) = Z (0) (T, V )d(p, q) dφdψµ(φ, ψ)χ∗(p,q) Z (1) (T, V ) (13.89)

This allows us to find Z(p,q) for different colour representations or colour projec-
tions. As the colour singlet is of special interest, we demand that the QGP be colour
singlet globally. For it we obtain,

Z(0,0) (T, V ) = Z (0) (T, V )Z (1) (T, V ) (13.90)


where
Z Z
Z (1) (T, V ) = dφdψµ(φ, ψ)Z (1) (φ, ψ, T, V ) (13.91)

Here Z (1) (T, V ) gives the correction term in the energy due to imposition of the
colour singlet condition. These integrations we perform numerically.

————————————————–
Problem 13.5: Find the range of the angles αq and αg .
484 Group Theory in Particle, Nuclear, and Hadron Physics

————————————————–

Once we obtain Z (0) (T, V ) and Z (1) (T, V ), we can obtain energy expression as,


E(p, q) = T 2 lnZ(p,q)
∂T
For the singlet case,

∂ ∂
E (0) = T 2 lnZ (0) (T, V ) and E (1) = T 2 lnZ (1) (T, V )
∂T ∂T
Thus E(0,0) = E (0) + E (1) . And we define,

E(0,0) E (1)
Def f = = 1 + (0)
E(0) E
The quantity Def f defined here describes the deviation of the QGP gas from the
Stefan-Boltzmann ideal gas behaviour. Thus it may be treated as a measure of the
effective number of degrees of freedom.
Let us now calculate E (0) and E (1) for the colour singlet projections case. From
eqn. (86),

37π 2 V T 3
Z (0) (T, V ) = exp( )
90
So from above,
37 2
π V T4
E0 =
30
This is the unprojected energy (i.e. with no colour restrictions whatsoever).
Next from eqn. (91),

π2 V T 3
Z (1) (φ, ψ, T, V ) = exp( A)
12
where
2 4
X 1 αq αq X 1 |αg | 1 |αg |
A = dq ( ( )4 − ( )2 ) + dg (− + ( − 1) − ( − 1) )
2 π π g=µ,ν,ρ,σ
2 π 2 π
q=r,g,b

————————————————–
Problem 13.6: Simplify to express the term A above in terms of angles φ and ψ
(These will be applicable to performing integrations over these angles for obtaining
Z (1) (T, V )).
————————————————–

Thus we get the Z (1) (φ, ψ, T, V ) term from which we obtain Z (1) (φ, ψ) by inte-
gration using Equation 13.91 and finally E (1) . Having obtained E (0) and E (1) , we
get Def f for colour the singlet projection case:

15 Y
Def f = 1 + ( ) (13.92)
74 Z
where
Quark Gluon Plasma(QGP) 485

1.0

0.9

0.8

0.7

o.a
D sing 161:
EFF
0.5
8-plet

2:7-F»l«st
0.4

0.3

0.2

0.1

o 1.0 2.0 3.0 4.O


1/3
< T V > /n c

FIGURE 13.3: Def f for singlet, octet, and 27-plet representation for two
flavours

Z π Z π
8
Y = d(φ/2)d(ψ/3)(sin 1/2(φ/2 + 3ψ/3) sin(φ/2)
−π −π 3π 2
π2 V T 3 A
sin 1/2(−φ/2 + 3ψ/3))2 A exp( )
12
Z π Z π
8
Z= 2
d(φ/2)d(ψ/3)(sin 1/2(φ/2 + 3ψ/3) sin(φ/2)
−π −π 3π

π2 V T 3 A
sin 1/2(−φ/2 + 3ψ/3))2 exp( )
12
and A is given in the solution of the above problem.
Note that as we have an equal number of quarks and anti-quarks in our calcu-
lations here, the chemical potential µ = 0. We may plot the case for colour singlet
projection [ [136], [137] ] This of course yields the putative global colour singlet state.
But we also want to know as to what is the behaviour of QGP for colour projec-
tions like the octet (1, 1) and the 27-plet (2,2). Do they escape to infinite energies
so that they have no low-energy manifestations? We therefore also project out the
octet and 27-plet representations for the quark gluon plasma [127]. First we need to
numerically evaluate Z(1,1) and Z(2,2) from Equation 13.70 where we already know
the character χ(1,1) and χ(2,2) . Next the corresponding Def f for that representation.
We give schematic plot Def f for singlet, octet, and 27-plet representation for
2-flavours in Figure 13.3. (Similar plots for zero-flavour (gluons) and 3-flavours, see
[127]).
486 Group Theory in Particle, Nuclear, and Hadron Physics

First let us concentrate upon the gluon singlet case [136], [137]. We find that for
large T V /(~c) > 2 the colour singlet does not play any significant role vis-a-vis the
colour uncorrelated states (i.e., the Stefan-Boltzmann colour unprojected states).
Meanwhile it is significant for finite sizes when T V /(~c) > 2 is small, say less than
two. Thus there occurs a rapid decrease in the number of internal (colour) degrees
of freedom of the QGP. One may call this a phase transition of the system. In
nuclear collisions relatively small plasma droplets will be formed, just of the order of
R ∼ 2.5f m, and our calculations demonstraate the relevance of colour confinement
effects for such experiments [137]. Indeed, this is along expected lines in QGP.

————————————————–
Problem 13.7: We found that the maximum colour correction took place ( i.e.,
E 1 is minimum ) at T V /hc ∼ 1.4. Given T = 160M eV what is R?
————————————————–
Problem 13.8: From Figure 13.3 for the singlet case, if we take the sharpest
1/3
drop at T V~c ∼ 1.5, then tabulate R(fm) and T(MeV) for R=0.1 fm to 1 km for a
couple of different values of R.
————————————————–

However a perplexing puzzle arises when we observe the plots for the octet and
27-plet representations. We find that for small T V /~c values the octet and the 27-
plet energies shoot up. Thus for the µ = 0 case there is a clear distinction between
the global colour singlet states, the global colour octet case and the global 27-plot
state. Clearly the octet and the 27-colour-plet states are moving to infinity and are
hence inaccessible to low-energy ground states, while the singlet states dominate
at low energies. This situation is also independent of the numbers of 0-, 2- and
3-flavours. The fact that the colour singlet representation gets favoured over the
octet, the 27-plet and other representations at low temperatures and smaller sizes,
supports the SU (3)c global colour symmetry concept.
However, at higher values of T V /~c all the unprojected states, singlet, octet, and
27-plet – become degenerate. There is nothing to distinguish one from the other. In
fact higher representations, take (3,3) of dimension 64, (4,4) of dimension 125, (5,5)
of dimension 216 and so forth, all of which will blow up at shorter T V /~c scales and
merge with all the others for higher values of the same. In fact we can extrapolate
to an infinite dimensional representation, and the conclusion remains the same.

————————————————–
Problem 13.9: Confirm the dimensions of the representation of (3,3),(4,4) and
(5,5) by using the Young Diagram technique.
————————————————–

Thus we conclude that all states – colour unprojected, singlet, octet, 27-plet,
64-plet up to infinite-colour-projection – with self-conjugate real representations of
µ = 0 QGP are degenerate in energy for higher values of T V 3 /~c. There is nothing
that distinguishes one state from the other. What does it mean?
What this connotes is that above say T V 3 /~c > 2 the QGP should have all the
colour projected states, going right up to infinity. It consists of the complete infinite-
dimensional Hilbert space. If colour singlet is the smallest dimensional (or minimal)
Quark Gluon Plasma(QGP) 487

colour-projected state, then this infinite-dimensional colour-projected QGP state is


the maximal colour projected state. Just as the colour singlet state is invariant and
unique in hadron physics, so is the infinite-dimensional maximal colour projection
state. Also it is unique in as much as nothing can be added to it. Hence it is invariant,
as it cannot be changed in this colour space.
So what our global colour projection technique for the full SU (3)c colour group
demonstrates, is that this QGP should be colour-invariant as an infinitely coloured
state. As this infinite and maximal colour projection holds uniformly for each point
in the QGP, it is a fully homogeneous and isotropic system in colour space. The situ-
ation is similar to the isotropy and homogeneity of 3-dimensional space in cosmology.
Thus in QGP only two forms of motion are possible in this homogeneous-isotropic
space – uniform expansion and uniform contraction. This is so as if any other mo-
tion were to occur, then some part of the QGP would look different with respect to
the other regions. This is forbidden by the infinitely and uniformly fully coloured
isotropic and homogeneous QGP.
How about the complete absorption of high-momentum protons in jet-
quenching? Clearly as our QGP is infinitely coloured, no coloured entity can pass
through it and complete absorption in the QGP fluid shall occur. Thus this previ-
ously vexing problem is also resolved in this model. Consequently, QGP is indeed a
“perfect fluid”.
But as per the QGP conundrum there is an empirical free gas nature of same
fluid QGP. How do we understand this puzzle?
For this let us look at the global nature of the character of a particular rep-
resentation χ(p,q) in the colour group SU (3)c . Let us now analyse the expression
for the character χα (p,q) . Note that it is a mathematical property of the character of
a representation, that once constructed it forgets as to how and what microscopic
entities created it. This is a well-known mathematical fact of group theory.
The important property of the definition of the group character for the invariant
group function is that it is independent of the microscopic structure of the states
(|α, β > in Equation 13.66) which transform under the irreducible representation α.
In other words, it does not matter how many particles the multiplet is made of.
Thus as we build up the colour singlet state at small T V 3 /~c values from current
quarks, anti-quarks and gluons, the contributing terms of 3 × 3 → 1 + . . . and
8 × 8 → 1 + . . . are obtained. But for higher > 2 values of T V 3 /~c where it is
degenerate with 8-plet, 27-plet, 64-plet till infinity of colour states, all these shall
clearly contribute for that particular singlet state existing as high T V 3 /~c values.
This reality arises due to the above fact that this singlet state accepts contributions
from any colour state which is available to create it at a particular energy.
Hence the quarks which produced the 3 ⊗ 3 → 1 + . . . singlet states are current
no more. They have been fattened out by all the infinite colour states such as 3 ⊗
(27 ⊗ 27) → 3 ⊗ (1 + . . . ) → 3. These are now quasi-particles or constituent-quarks.
Next as the temperature of the QGP falls, all the highly coloured representations
– 264-, 125-, 64-, 27- plet and octet – fall out of contention by disappearing to higher
energies, one by one. Likewise, the colour singlet state constituent quarks will move
down.
Note that the colour singlet states which have moved down in Figure 13.3 con-
sist only of constituent or quasi-quarks with a 3 × 3 → 1 + . . . representation.
Also there are no gluons or colour octet states, as these too had disappeared to
infinity. Thus this colour singlet state in the phase transformation is a free quark-
488 Group Theory in Particle, Nuclear, and Hadron Physics

antiquark state, but now of the constituent kind. When these lead to the creation of
baryon-antibaryon states and meson states, it is the manifestation of the pure quark
character in these sates. And hence we see that the quark scaling as discovered by
the PHENIX group, which we have discussed earlier, is a manifestation of the above
quark nature of the colour singlet in the QGP phase transformation. Thus this model
is able to produce a resolution of the QGP conundrum – is it a liquid or a gas? The
answer here is that it is both, arising from the quark (and gluon) degrees of freedom.
These provide the QGP with simultaneous liquid-like and gaseous characteristics!

13.7 Solutions of Problems

Solution 13.1:
These can be checked against the dj -representation of SO(3) in quantum me-
chanics,
 
cos α/2 − sin α/2
d1/2 (α) =
sin α/2 cos α/2
and
 √ 
(1 + cos α)/2
√ −(sin α)/ 2 (1 − cos α)/2

1
d (α) =  (sin α)/ 2 cos α √ (sin α)/ 2 
(1 − cos α)/2 −(sin α)/ 2 (1 + cos α)/2

whence χ1/2 (α) = 2 cos α/2 and χ1 (α) = 1 + 2 cos α


————————————————–
Solution 13.2:
The Young Diagram for the anti-triplet for SU(3) is:

3̄ = (0, 1, 0) = (h13 − h23 , h23 − h33 , 0), and so h11 = 1, h23 = 1, h33 = 0

1
X 0
e−iψ(h22 +1) . sin[((2 − h022 )/2)φ](sin(φ/2))−1 ))
(0,1,0)
χSU (3) = e4iψ/3 (
h022 =0

= e4iψ/3 (e−iψ .2 sin(φ/2) cos(φ/2)(sin(φ/2))−1 + e−2iψ )


= e−i/3(θ11 +θ22 −2θ33 ) (1 + ei(θ11 −θ22 ) + ei(θ22 −θ33 ) )
= eiθ11 + eiθ22 + eiθ33

(0,1,0)
χSU (3) = Σeiθii
Quark Gluon Plasma(QGP) 489
(0,1,0) (1,0,0)
Note χSU (3) = (χSU (3) )∗
————————————————–
Solution 13.3:
The Young Diagram for the 27-Plot is

(2, 2, 0) = (h13 − h23 , h23 − h33 , 0), h13 = 4, h23 = 2, h33 = 0.


Checking the dimensions of the Young Diagram,
3×4×5×6×2×3
D= = 27
2×4×5×2

4 2
X X 0 0
e−iψ(h12 +h22 ) · sin((h012 + 1 − h022 )/2)φ)(sin(φ/2))−1 )
(2,2,0)
χSU (3) = e4iψ (
h012 =2 h022 =0

4
X 0
= e4iψ ( (e−iψh22 ) sin((h012 + 1)/2)φ)(sin(φ/2))−1 )
h012 =2
0
+(e−iψ(h12 +1) ) sin((h012 )/2)φ)(sin(φ/2))−1 )
0
+(e−iψ(h22 +1) ) sin((h012 − 1)/2)φ)(sin(φ/2))−1 ))
= e4iψ (e−2iψ (sin(3φ/2))(sin(φ/2))−1 )
+e−3iψ (sin(2φ/2))(sin(φ/2))−1 )
+e−4iψ (sin(φ/2))(sin(φ/2))−1 )
+e−3iψ (sin(4φ/2))(sin(φ/2))−1 )
+e−4iψ (sin(3φ/2))(sin(φ/2))−1 )
+e−5iψ (sin(2φ/2))(sin(φ/2))−1 )
+e−4iψ (sin(5φ/2))(sin(φ/2))−1 )
+e−5iψ (sin(4φ/2))(sin(φ/2))−1 )
+e−6iψ (sin(3φ/2))(sin(φ/2))−1 ))

χSU (3) = 2 + 2 cos φ + 2 cos(3φ/2) cos(φ/2) + (eiψ + e−iψ )(2 cos(φ/2)


(2,2,0)

+4 cos(φ/2) cos φ) + [1 + 2 cos φ][e2iψ + e−2iψ + cos 2φ/2]


Finally,

(2,2,0)
χSU (3) = 2 + 2[cos φ + cos(3φ/2) cos φ/2] + 2(1 + 2 cos φ)[cos(φ/2 + ψ)

+ cos(−φ/2 + ψ) + cos 2ψ + (1/2) cos φ]


————————————————–
490 Group Theory in Particle, Nuclear, and Hadron Physics

Solution 13.4:

Z π Z π
V = 64 d(φ/2)d(ψ/3) sin2 φ/2 sin2 (1/2(φ/2 + ψ)) sin2 (1/2(−φ/2 + ψ))
−π −π
Z π Z π
= 64 d(φ/2)d(ψ/3) sin2 φ/2(cos φ/2 − cos ψ)
−π −π
Z π Z π
= 16 dxdy sin2 x(cos x − cos 3y)2
−π −π
Z π
= 16 dx(2π(sin2 x cos2 x) + sin2 x(π + 0))
−π

= 16(2π(0 + π/4) + π(0 + π))


= 24π 2
————————————————–
Solution 13.5:
Now, −π ≤ φ/2 ≤ π and −π ≤ ψ/3 ≤ π. Then,
   
αr φ/2 + ψ/3
αq = αg  = −φ/2 + ψ/3 → −2π ≤ αq ≤ 2π
αb −2ψ/3

     
αµ φ −2π ≤ αµ ≤ 2π
αν  −φ/2 + ψ  −4π ≤ αν ≤ 4π 
 αρ  =  φ/2 + ψ  →  −4π ≤ αρ ≤ 4π 
αg =      

ασ 0 ασ = 0
————————————————–
Solution 13.6:

2 4
X 1 αq αq X 1 |αg | 1 |αg |
A = dq ( ( )4 − ( )2 ) + dg (− + ( − 1) − ( − 1) )
2 π π g=µ,ν,ρ,σ
2 π 2 π
q=r,g,b

Where using expression for α’s from Equations 13.74, 13.75, and 13.76 in terms
of angle (φ/2) and ψ/3 (as integration in Z shall be done over these values) we obtain
1 1 1
A = 4( (φ/2 + ψ/3)4 − 2 (φ/2 + ψ/3)2 + 4 (−φ/2 + ψ/3)4
2π 4 π 2π
1 1 1
− 2 (−φ/2 + ψ/3)2 + 4 (−2ψ/3)4 − 2 (−2ψ/3)2 )
π 2π π
2 1 2 | − φ/2 + 3ψ |
+2(−3/2 + ( |φ/2| − 1)2 − ( |φ/2| − 1)4 + ( 3
− 1)2
π 2 π π

1 | − φ/2 + 3 | |φ/2 + 3ψ | 3ψ
1 |φ/2 + 3 |
− ( − 1)4 + ( 3
− 1)2 − ( − 1)4 )
2 π π 2 π
Next expand and simplify.
Quark Gluon Plasma(QGP) 491

————————————————–
Solution 13.7:

T V 1/3 4π 3
= 1.4 , V = R ,
~c 3
1.4 ~c
R= 1/3
= 1.08f m (f or ~c = 200M eV − f m)
T ( 4π
3
)
————————————————–
Solution 13.8:
T V 1/3
For ~c
∼ 1.5
R(f m) 0.1 1.0 5.0 1013 (1cm) 1018 (1km)
T (M eV ) 1861 186 37 1.8 × 10−11 1.8 × 10−16
————————————————–
Solution 13.9:
(3,3) is

3×4×5×6×7×8×2×3×4
→ 2×3×5×6×7×2×3
= 64

3×4×5×6×7×8×9×10×2×3×4×5
(4,4) is → 2×3×4×6×7×8×9×2×3×4
= 125

(5,5) is → · · · = 216

Note for these self-conjugate (real) representations given as (p,p), the dimension
is simply (p + 1)3 .
————————————————–
Chapter 14
Topology for Hadrons

14.1 Why Topology? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 493


14.2 Homotopy Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 495
14.3 Linear and Non-Linear Sigma Models . . . . . . . . . . . . . . . . . . . . . . . . . . . 498
14.4 Skyrme Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 500
14.5 SU(3) Adjoint Representation Skyrme Model . . . . . . . . . . . . . . . . . . . 503
14.6 Appendix G: Introduction to Topology . . . . . . . . . . . . . . . . . . . . . . . . . . 505
14.7 Solutions of Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513

14.1 Why Topology?


In order to understand the applicability of topology to hadron science, it is
imperative to first comprehend the distinctions which differentiate it from its close
cousin, geometry.
Now, geometry –as it deals with fixed shapes which permit their easy and con-
crete illustration via graphical diagrams – is intuitive in nature. Hence geometry
education represents an early component of student education. On the contrary,
topology – as it involves more abstract depictions and fluid structures which often
defy simplistic visualizational reductions in 2-dimensional space – is non-intuitive
in nature. Thus topology education is imparted much later on in life. As such one
may view geometry and topology as providing complementary depictions about re-
ality that we study as physicists (Appendix G). However, knowledge in its wholeness
should encompass both the intuitive and the non-intuitive components of our phys-
ical reality.
In topology the most primitive object, a collection of points, is denominated a
differential manifold. In order to define the simplest topological structure, a condition
of smoothness is imposed on it, so as to ensure a modicum of continuity. These
topological structures and transformations are independent of the concept of length
or a metric. Thereafter one introduces a new property on the manifold, that of an
affine connection, plus a new structure which brings in a notion of length through
a metric. In general the concepts of the affine connection and that of the metric are
quite independent of one other. However, as it turns out, what is physically relevant
is that an affine connection is defined in terms of the metric – the so-called metric
connection.
One should note that in this metric structure, the topological framework is any-
way simultaneously present in the background, implying that both the geometric
and topological structures are present intrinsically.

493
494 Group Theory in Particle, Nuclear, and Hadron Physics

From the special theory of relativity and the general theory of relativity we have
learned that the metric or length defines all that can be learned about the structure
of the space-time. If we know the metric, we have learned as much as possible about
the fabric of space-time. But one should not forget about the intrinsic significance of
the “non-metric”. It should have a certtain physical significance as well. In fact, this
corresponds to the topological structure. Invariants of the topology provide further
basic information about the space-time structure. Thus topology plays an important
part in the study of the structure of space-time [138].
Hence geometry – which is based on properties dependent upon the existence of
a metric – and topology – which is founded on characteristics independent of length
– play a complementary role here. Not only does it appear that the topological struc-
tures and the geometrical structures are complementary aspects of the mathematical
space, in as much as a metric may be present or be absent – the complementarity
is exclusive. Moreover, in as much as these are the only two possibilities which a
physically relevant structure can possess – the complementarity is exhaustive too. A
philosophical analogy would be the complementary self-other dualism – as opposed
to the antagonistic good-evil dualism – envisioned in Manichaeanism.
Viewed in this manner, these two frameworks constitute a fundamental structural
duality of space-time. One should note that this duality is of a generic nature,
not dependent upon the details of either geometry or topology involved. We shall
term this new fundamental duality as the geometry-topology complementarity
paradigm [138]. This complementarity and the dual aspect of the geometric and
the topological structure hypothesised as the geometry-topology complementarity
paradigm, holds for the structure of 4-dimensional space-time. However how does
this translate into the symmetries of our physical theories of the particles, the nucleus
and the hadrons? We shall discuss this below.
Now, all the conservation laws forming the basis of the classical field theories, the
non-relativistic field theories and the relativistic field theories, arise as a consequence
of Noether’s theorem. As to space-time and internal symmetries, the Noether con-
servation law in physics is a consequence of the invariance of the Lagrangian under
a continuous symmetry transformation. The resulting Noether charges correspond
to field theories with a single physical vacuum. This is essentially the basis of most
of the applications of group theory in what we have accomplished in the previous
chapters.
The above would correspond broadly to the “geometric” equivalence of the phys-
ical theories. The other complementary concept of topology, is what we talk now.
A solution for a wave equation that is non-dispersive is denominated a solitary
wave. Hence we define a ‘soliton’ as any extended but spatially localised and non-
dispersive solution of a non-linear wave equation [31], [139].
Topological solitons (in contrast to non-topological solitons, which we do not
discuss here) are finite energy non-dissipative solutions of the Lagrangian equations
for which the boundary conditions at spatial infinity are topologically different from
those valid for the single physical vacuum [139]. These topological soliton states
are classified by specifying topologically inequivalent boundary conditions at infin-
ity, and invoking selection rules which rule out transitions from one class of soliton
states to another or to the vacuum. These are called the “topological conservation
laws”. Unlike Noether Conservation Laws, the topological conservation laws do not
arise from some symmetry of the Lagrangian but are a consequence of the possibility
of defining topologically distinct classes of finite energy and deriving non-singular
Topology for Hadrons 495

solutions of the Lagrange equations. Thus these are classes of soliton solutions which
cannot be topologically deformed into one other. To obtain these topological con-
servation laws it is necessary to define a homotopic group language.
The Lie algebra defines the local properties of a Lie group. But global properties
such as connectedness (see Appendix G) are not uniquely specified by the commu-
tator algebra. The group manifold, specified by the group parameters, defines the
topological characteristics of the group. Hence the matrix representation of a group
G, the angle variables ϕα parametrise the group V = expl − (φα Tα ), The space {φα }
over which φα varies is defined as the group manifold G. To ensure one-to-one cor-
respondence between the group elements and φα , these are accordingly constrained.
A group is called compact if its manifold is compact. The word compact here means
that the entity is of finite size or volume. The topology of group manifolds plays a
basic role in gauge field theories, especially as providing complementary topological
conservation laws to the geometrical Noether conservation laws.
It should be emphasised here that the perturbative aspects of the quantum
field theories as providing us with the geometrical Noether conservation laws are
independent and complementary to the non-perturbative aspect of the quantum
field theory and are manifested by topological conservation laws. Both are needed to
simultaneously provide a complete and consistent description of the whole physical
reality embodied in the quantum field theory.
The complementarity and the dual aspect of the geometric and the topolog-
ical structures, as hypothesized above as the geometry-topology complementarity
paradigm, holds for the structure of 4-dimensional space-time. However this does
not seem to translate into the symmetries of our physical theories of the particles,
the nucleus and the hadrons. Why not? The reason has to do with the geometrical
nature of the Noether current in gauge theories. So to account for the duality as
to the understanding of the particles, we invoke the global-local complementarity
paradigm, discussed in Appendix G. This explains the use of both the global and
the local aspects to understand the basic nature of the particles.

14.2 Homotopy Groups


For a group manifold, connectedness is a property basic to defining the homotopy
group – which is a group of connected paths on the group manifold. First to get an
intuitive idea of the concept of a path. We define two paths on a group manifold
with the same end points as topologically equivalent if there exists a function which
can continuously deform one into the other. See Figure 14.1 (a) (LHS). Homotopy
is the continuous function which deforms one path into another and these paths are
termed homotopic. So given the beginning and the end as two points, homotopy is
an equivalence relation for a collection of paths between these two points. See Figure
14.1 (a) (RHS) for paths α and β as belonging to different homotopy classes. The
number of classes of paths on surface (RHS) is infinite. A path that wraps around
the ring once cannot be distorted into another path which wraps around it twice
and so on.
Let a path x in a space X be defined as a continuous function x(r) of a real
496 Group Theory in Particle, Nuclear, and Hadron Physics

parameter r, such that in the interval (0 ≤ r ≤ 1) x(r) is a point in space X. If a


path x connects the points A and B then x(0) = A, x(1) = B. We define a loop (or
a closed path) at A when e.g. x(0) = x(1) = A. Consider two closed paths x(r) and
y(r) with a function f (s, r) such that f (o, r) = x(r), f (1, r) = y(r). Then x and y
are homotopic which is written as x ∼ y.
The inverse of path x is x−1 and is defined as,

(x−1 ) = x(1 − r) (14.1)


which means that it is an oppositely transversed path. If the path-x end-point con-
nects with the beginning point of path-y, i.e., x(1) = y(0), we define the product
path c = ab as:
1
z (r) = x(2r) f or 0 ≤ r ≤
2
1
z (r) = y(2r − 1) f or ≤r≤1 (14.2)
2
−1
We define a null path as a single point. If x ∼ y then xy is homotopic to the
above null path.
Let us define the class of paths homotopic to x as [x]. These must have the same
end points.
We define the multiplication of these homotopy classes as a product,

[x][y] = [xy] (14.3)


We show below that this multiplication satisfies the group laws.
(a) Closure: If [x], [y] ∈ G then [x][y] = [xy] ∈ G
(b) Associativity: ([x][y])[z] = [x]([y][z])
(c) Identity: null path [1] class. [x][1] = [x]
(d) Inverse: [x−1 ][x]=[x−1 x] = [1] so [x]−1 = [x−1 ]
This group G is called the fundamental group or the first homotopy group
of space X and specified as Π1 (X) [65]. In addition to the compactness of a group
manifold we require the concept of connectedness (Appendix G). If every closed path
in a manifold of a connected group is deformed continuously to a point, then the
group is called “simply connected”. Otherwise it is said to be multiply connected.
We now look at compact manifolds defined on spheres denoted by S n where n
is the number of dimensions on the surface of the sphere. So S 0 is a point, S 1 is
a circle, S 2 is the usual sphere in the 3-dimensional space and S 3 is a sphere in
4-dimensional space.
Let us define the nth homotopy group as Πn (X) where the n-sphere S n maps
onto the topological space X. Hence the mapping is from a point in S n to a point in
X. Take Π0 (S)1 . Hence take a point outside a circle S 1 [Figure-1(b)]. Now this point
will map onto a single point on S 1 . Topologically distinct paths are obtained by going
around the point on S 1 (which is any point at all) clockwise, with a winding number
+1 or anticlockwise with a winding number −1. We can go around m-number of
times and these are topologically distinct paths and thus are specified by integer Z
winding number and can be infinitely connected. For Π0 (S 2 ), as it is still a single
point on S 2 we have Z winding number here too. Thus,
 
Π0 S k = Z (14.4)
Topology for Hadrons 497

1 (1,2) a

(a)

P
* (1,2)
2

(b)

0 +1 -l -2

i j
(c)

a
b

a-space (5-space

FIGURE 14.1: Topological mapping and winding numbers

Clearly Π0 (S k ) really brings out the connectedness property of a group manifold.


Now let us study the mapping S 1 → S 1 , i.e., the homotopy group Π1 (S 1 ). It is
just the mapping of a circle to another circle (Figure 14.1 (c)) with the point ‘a’ on
circle ‘i’ in α-space mapping and point ‘b’ on circle j in β-space. A mapping when
all the points on ‘i’ are mapped on to the point ‘b’ on j. As all the points on the
arc in circle j can be continuously deformed to the point b it gives the unit element
of the homotopy group with winding number zero. Let us tie a string around the
circle ‘j’ and this cannot be continuously deformed into the point ‘b’ and hence the
mapping has a winding number 1. Similarly a string wound around ‘j’ has winding
number 2. Thus it is infinitely connected mapping.

Π1 S 1 = Z

(14.5)
2 2

Next Π1 S where S dimensional surface of a solid sphere in 3-dimensions.
But now mapping S 1 → S 2 leads to any string covering the surface to slip off and
collapse topologically to a point. So to say, “You cannot lasso a basketball”. The
topology of S 2 is trivial as it is simply connected. This is true of any S n (n ≥ 2).
And so

Π1 (S n ) = 0 f or n > 1 (14.6)
n m
We can generalize to a homotopy group for mapping S →S (m, n integer)
which is designated as Πn (S m ),

Πn (S n ) = Z
Πn (S m ) = 0 , n < m (except f or n = 0) (14.7)
498 Group Theory in Particle, Nuclear, and Hadron Physics

Thus mapping S n → S m is trivial for n < m, For n = m the mapping is non-


trivial and a topological invariant can be defined related to the integer winding
number which is how many times a sphere is wrapped around the other sphere.
The U (1) group is defined with elements of the form

U = eiθ (14.8)
1
The manifold of U (1) (or of SO(2)) is the unit circle S which is compact but
not simply connected. Thus,

Π1 (U (1)) = Π1 S 1 = Z

(14.9)
Thus there are infinitely many classes of closed paths which cannot be contin-
uously deformed into each other (for a simply connected space X every path is
homotopic to the null path Π1 (x) = 1). This infinite connectivity of U (1) is impor-
tant as Quantum Electrodynamics exhibits the group structure U (1)em and thus it
allows for the quantum Bohm-Aharonov effect to be possible [46].
Let us parametrize the SU (2) group as

iτ a φa (σ + i~τ · ~π )
U = exp( )= (14.10)
fπ fπ
In term of a scalar σ and a pseudoscalar ~π , with the standard notation for ~τ we
find:
 
1 σ+ıπ 3 π2 +iπ1
U= (14.11)
fπ −π 2 +iπ 1 σ−iπ3
Subject to the condition that

σ 2 +π12 +π22 +π33 =fπ2 (14.12)


3
for SU (2) to hold. Now the equation of the S surface on a four-dimensional space
is given by Equation 14.11. So SU (2) is defined by a group manifold of S 3 .
So using Equation 14.7,

Π1 (SU (2)) = Π1 S 3 = 0

(14.13)
These will be utilized in studies on the topological structure of certain special
Lagrangians as we discuss below.

14.3 Linear and Non-Linear Sigma Models


The Linear sigma model starts by taking the fields – an isotriplet of pseudoscalar
pions ~π =(π1, π2, π3 ), an isosinglet scalar and a σ meson, and doublet of massless
nucleons N = (p, n). If we take the masses of these mesons and σ-meson as equal,
then we get a global symmetry of O(4). O(4) has 4(4−1) 2
= 6 generators. We may
break the O(4) symmetry spontaneously to O(3) (with its three-generators). But as
O(4) is locally isomorphic to SU (2)L ⊗SU (2)R one prefers to take it (with its 3+3=6
Topology for Hadrons 499

generators) and break it to SU (2)L+R (with its three generators) in the presence of
massless nucleon N .
Let us write the linear sigma model Lagrangians as

1
(∂µ σ)2 +(∂µ ~π )2 +N̄ iγ µ γµ N + g N̄ (σ + i~τ · ~π γ5 ) N − V σ 2 +π 2
 
L= (14.14)
2
When the potential V σ 2 +π 2 has the usual Englert-Brout or Higgs quartic


form [46]

 µ 2 2 2  λ 2 2 2
V σ 2 +π 2 = σ +π + σ +π (14.15)
2 4
with µ and λ constants. The choice of V σ 2 +π 2 having quartic terms in σ and π,


arises from the fact that the corresponding quantum field is renormalisable.
This Lagrangian is invariant under a pair of SU (2) ‘vector’ and ‘axial-vector’
transformations.
The global chiral SU (2)L ⊗ SU (2)R group is spontaneously broken in the usual
manner [46]. The Englert-Brout/Higgs potential V (σ 2 + π 2 ) has a minimum at
r
−µ2
σ 2 + π 2 = v 2 with v = (14.16)
λ
Choosing < 0 | σ | 0 > = v (such that σ 0 = σ − v) produces standard equation
for the scalar part and the fermions acquire a mass mn = gv where g is the meson-
nucleon coupling constant and v is the displacement of the vacuum. In standard
manner the isotriplet ~π arises as the Nambu-Goldstone boson triplet with mπ = 0
1
whereas the scalar boson acquires a mass mσ ∼ λ 2 (for constant v).
Thus a light scalar meson is a clear prediction of the linear sigma model. What
is the experimental situation regarding this scalar meson? A look at the Particle
Data Group result shows the following scalar particles [53]:
(1) f0 (500) with m=(400-500) MeV and with width Γ = (400 − 700) MeV,
(2) f0 (980),
(3) f0 (1370),
(4) f0 (1500),
(5) f0 (1760.
Out of these the lightest f0 (500) may be identified with the sigma meson of the
linear sigma model. However there are doubts as to its very existence. So much so
that the Particle Data Group themselves state that, “the interpretation of this entity
as a particle is controversial” [53]. So if the σ-meson is so reluctant to show up in
the laboratory, how do we account for this reality?
Now above we have treated σ and ~π as independent fields and coupled to the
fermion field eqn. (15). One may however hide the σ-meson by imposing the con-
straints σ 2 +~π 2 =fπ 2 which is called the non-linear sigma model. In this model one
takes the attitude that σ meson is not observable in the laboratory as it is hidden
inside this above so-called chiral circle.
1
A better and more consistent method is to note that as mσ ∼λ 2 we can make
mσ → ∞ by setting λ arbitrarily large. With the further dropping of the N -term
in the non-linear sigma model, we shall obtain the Skyrme Model, which has a
unique topological structure which is identified with the nucleon N . Note that σ
is actually as real as the ~π field. σ-meson as such is akin to the Higgs boson of
500 Group Theory in Particle, Nuclear, and Hadron Physics

the Standard Model. Actually the missing σ-meson is called the ”Higgs boson of
the strong interaction” [PDG]. So there are actually two Higgs bosons – one of the
Standard Model and the other of the strong interaction. Now we are advocating to
take mσ → ∞ to obtain the Skyrme Model. What does it mean to take mσ → ∞
? As far as we know there are no known physical particles whose mass is infinity.
Actually an infinite mass means that the particle disappears out of contention and
physically is not relevant.
Note that we take the physical point mass ∼ 140 MeV to be sufficiently small
compared to the hadronic/baryonic mass of ∼ 1 GeV to be able to treat mπ ∼ 0 as a
sufficiently good approximation to understand the low energy structure of hadrons.
In the opposite end now, we take mσ → ∞ to obtain the fundamental structure
manifested in the Skyrme model, as a good approximation for a sufficiently heavy
sigma meson mass. Thus instead of making it of infinite mass we treat it as being of
sufficiently heavy mass that it does not mess around the mass of the lighter baryons
or hadrons. Hence the existence of the fundamental theory of the Skyrme model
predicts a very heavy sigma meson. So the missing σ-meson of the strong interaction
is actually absent at low energies simply because it is too heavy. We cannot supply
a number, but given the present searches for the Higgs boson at CERN it should
be about or more heavy than 125 GeV. This sigma scalar meson however partaking
only in the strong interaction will have no (τ + τ − ) decay-channel, in contrast to the
electro-Weak Higgs boson. In as much as controversy still exists about this channel,
and as the last word on this has not yet been spoken by the experimentalists (as
of beginning of 2016), this 125 GeV scalar meson as of now may very well be this
missing sigma meson [140]. So whether an electro-weak Higgs boson exists or not,
independently of this, the topological Skyrms model predicts a very heavy scalar
meson of strong interaction. Experimentalist should continue searching for it.

14.4 Skyrme Model


In the linear sigma model it is not essential to include the fermionic-nucleonic
field N . We may very well write the linear sigma model without the N term and
reduce Equation 14.14 to:
1
(∂µ σ)2 +(∂µ ~π )2 − V σ 2 +π 2
 
L= (14.17)
2
where V(σ 2 +π 2 ) is given by Equation 14.15. One can show that [31] this Lagrangian
is invariant under vector and axial-vector transformations and satisfies the standard
SU (2)L ⊗ SU (2)R algebra. Because µ2 <0, so spontaneous symmetry breaking still
occurs with the minimum of the potential at:
r
−µ2
σ 2 + π 2 = f 2 with f = (14.18)
λ
Making the standard choices < 0 | σ | 0 > = f and < 0 | π | 0 > = 0 and putting

the shifted field (σ 0 = σ − f ) with Equation 14.18 we obtain mσ = 2|µ| and mπ
vanishes as Nambu-Goldstone boson.
Topology for Hadrons 501

Now we wish to eliminate the σ-field from above. We may do so by taking the
limit mσ → ∞ (or more physically making it as of very large magnitude) and
placing ~π in a non-linear representation of the SU (2)I group. Let us break the chiral
symmetry in the linear sigma model in the form (see problem below),

1   λ
L= T r ∂µ φ† ∂µ φ − (T rφ† φ−2f 2 )2 (14.19)
2 4
φ = σ+i~τ · ~π . We wish to study the low energy hadron phenomenon and so the
mσ → ∞ is a reasonable physically realisable option to give the physics relevant to

us. We may adjust mσ by making a proper choice of µ. As mσ ∼ λ we can study
this limit by considering as to what happens to Equation 14.19 as λ → ∞.
Let us impose a constraint φ† φ=f 2 with a unitary matrix,

φ(x)
U (x) =; U †U − 1 (14.20)
f
Thus Equation 14.18 is transformed into a non-linear sigma model as:
1 h i
L= T r ∂µ U † ∂ µ U ; U † U = 1 (14.21)
2
The linear sigma model (Equation 14.19) is linear as the matrix φ sits in the
general linear group of Hermitian matrices. However the non-linear σ-model (eqn.
(21)) is non-linear as U (x) is an element of the group SU (2)L+R , i.e.,

U (x) = exp((2iT a π a /Fπ ), (a = 1, 2, 3) (14.22)

————————————————–
Problem 14.1: Demonstrate that without the interaction term, the Lagrangian
in Equation 14.17 can be written as in Equation 14.21.
————————————————–

It turns out that the Lagrangian of Equation 14.17 cannot support an acceptable
static solution as it would have zero energy and zero size. Hence Skyrme added a
quartic interaction so that the final Skyrme Lagrangian looks like.

f2 1
T r ∂ µ U + ∂µ U + T r[U + ∂µ U , U + ∂ν U ]2

LS = (14.23)
4 32a2
Now given Lµ = −iU + ∂µ U the Skyrme Lagrangian may be written as:

f2 1
LS = T r (Lµ Lµ ) − T r[Lµ , Lν ]2 (14.24)
4 32a2

————————————————–
Unsolved Problem 14.1: Prove Equation 14.24.
————————————————–

Let us look at the soliton structure present in the Lagrangian in Equation 14.24.
We surmise that the requirement of the soliton exhibiting finite energy demands
x, ~t) must approach a constant matrix U0 as r = |~
that the field U (~ x| → ∞. We know
502 Group Theory in Particle, Nuclear, and Hadron Physics

that the Lagrangian density LS is invariant under the global chiral transformation
and hence the constant matrix U0 can be reduced to the unit matrix 1 by a chiral
rotation without affecting LS . Therefore, without loss of any information, we can
in the most general manner set the boundary condition U → 1 as r → ∞. As for
U → 1 there is no angular dependence and hence for r → ∞ all the points at spatial
infinity may be identified as a single point. Thus this converts Euclidean space R3
of the coordinate space ~x for constant time to a 3-sphere S 3 . Now it can be shown
[31] that the spontaneous breaking of SU (2)L ⊗ SU (2)R yields the factor group:

G/H = M anif old of SU (2) (14.25)


3
We know that the topology of the SU (2) manifold is S . Thus the relevant
topological mapping of the Skyrme Lagrangian is:

S 3 (space) → S 3 (f ield) (14.26)


From the homotopy group we have,

Π3 (S 3 ) = Z (14.27)
where Z is the additive group of integers. Thus we expect here nontrivial topologies
characterised by this integer winding number m, related to topological conserved
currents and charges. Whence the topological baryon current is [31],
i µναβ
bµ =  T r Lν Lα Lβ (14.28)
24π 2
µ
as ∂µ b =R 0 and since U (~
x) is independent of time, the topological charge is defined
as B = d3 xb0 which is
Z
i ijk
B=  d 3 x T r L i Lj Lk (14.29)
24π 2
where baryon number is equal to the topological winding number m,

B=m (14.30)
as bµ is not a total divergence and therefore can be arbitrarily localised in space. As
bµ is a number current and therefore the topological baryonic charge is not coupled
to any long-range field. Thus as independent local field it counts simply additively
[65]. Note that the field U 0 maps all ~
x to the identity of SU(2) and thus B = 0 sector
consists of all maps which deform continuously to U 0 . Therefore the trivial B = 0
has no solitons and has baryon number zero. It thus corresponds to the standard
meson physics solution of the non-linear sigma model.
The most important conclusion of the Skyrme model is that the B = 1 solution
corresponds to a spin- 1/2 nucleon – which is a fermion arising topologically from
a Bose field. Note this an amazing fact. We are used to seeing bosons arising from
spin-1/2 entities as, for example,

→1

→ − → − →
2
+ 21 = 0 + 1 .
However the fermion arising from the Bose field is an interesting aspect of global
topological structure. If this structure is physically relevant and correct, then it
provides us with an important and basic piece of information. Successes of the gauge
theories in the last 50 years or so – on the foundations of which is built our consistent
SU (3)c ⊗ SU (2)L ⊗ U (1)Y Standard Model – has many scientists feel that the
Topology for Hadrons 503

entire physical reality (including what is not covered by the Standard Model) should
be a local gauge theory. This idea rejects any global symmetry and structures as
mere approximations to some underlying local gauge structure. However, the Skyrme
model demonstrates that, if it describes hadrons fundamentally in a consistent and
basic manner via a topological language, then the global symmetry structure is
fundamental and basic as well - at least as far as describing the physical reality as a
complete and monolithic whole. This seems to be demanding a duality in the local
and global symmetries in the microscopic and macroscopic structures of physical
reality. This duality bears a striking resemblance to the well-known wave-particle
duality in quantum mechanics.
Here we have suggested a new global-local complementarity paradigm to un-
derstand the above duality. More on the B = 1 solution which numerically can be
obtained in the “hedgehog form” (where isospin points in the radial direction):

U (r) = eiτ ·r̂θ(r) (14.31)


We demand the boundary condition as θ (r) =π at r = 0 to θ (r) → 0 at r =∞,
while ~
x covers the SU (2) manifold exactly once under the map eqn. (31). This is
giving us a unit winding number mapping. All mappings homotopic to it constitute
the B = 1 sector of the Skyrme model. We identify this B = 1 entity as the baryon
number to be associated with the baryons N . (B = −1 are topologically distinct
anti-baryons ) Thus the B = 1 topological charge be defined as :
 
1 1
B= θ (0) −θ (∞) − (Sin 2θ(0) − Sin(2θ (∞) ) (14.32)
λ 2
When the skyrmion is rotated adiabatically in the internal symmetry space it
generates a rotational band:
1
EJ=T =m0 + J(J + 1) (14.33)
2I
where m0 is the mass of the classical Skyrmion, I is the moment of inertia, T, J are
isospin and angular momentum with the equality condition. Thus,
1 3 5
T = J = , , ...... (14.34)
2 2 2
where at low energies we have N and ∆ states. As per the QCD analysis for arbitrary
Nc the baryon mass Mo ∼Nc while the second term may go as N1c . However the
static properties of N and 4 agree to within 30% of the experimental values.

14.5 SU(3) Adjoint Representation Skyrme Model


Note that in the above SU (2) Skyrme Model the defining U matrix was given
in terms of the adjoint representation of the SU (2) group as a traceless isovector.
 0 
π + π3 π1− iπ2 !

2
π √
2

2 1
πα β= 
π0
 = π1√
+iπ2 π3
= √ τ j πj (14.35)

π − 2 √ √ 2
2 2
504 Group Theory in Particle, Nuclear, and Hadron Physics

In SU (3) the adjoint representation can be given in terms of Cartesian coordi-


nates Pj (j = 1, 2 . . . 8) of the group SO(3) given as 3 × 3 matrix for SU (3),
−iP2 −iP5
 P P3 P1√ P4√

8
 √6 + √2
 2 2

1

P1√+iP2 P8 P3 P6√−iP7
√ λ j Pj = 2

6
− √
2 2
(14.36)
2  P4√+iP5 P6√+iP7 2

 
2 2

6
P 8

This is identified with the 0− meson octet as:

π0 π0
 
√ +√ π+ K+
√ 6

 2


0 0
λa π a (x, t) = 2 π − π
−√ +√π
K0 (14.37)
2 6
− ¯ 0 
 2


K K 0 − 6µ

which arise in the SU (3) field in the form:


 a a 
λ π (x.t)
U (x, t) = exp i (14.38)
g
where a = 1, 2, . . . 8.
This is the SU (3) field utilised in the Skyrme Lagrangian,

f2   1
LS = T r ∂µ U † ∂ µ U − T r[U + ∂µ U , U + ∂ν U ]2 (14.39)
4 32a2
The baryon number in Equation 14.29 still yields the SU (3) corresponding
baryon charge which is still quantised as per Equation 14.30. So in the SU (3) Skyrme
model with the adjoint representation of the 0− meson providing the basic field,
there is no fundamental structural change from the SU (2) Skyrme model. The same
topological soliton exists with a ground state baryon present [31].
So we take this fact to imply that, in the adjoint representation of SU (3) in
the Skyrme model, there exists a topological baryon number which is an extremely
important property. Remember in our discussion about the adjoint representation
of the eightfold way model, the fact is that the 1/2 baryon had no baryon number to
specify it. The 3 ⊗ 3 ⊗ 3 = 1 ⊕ 8 ⊕ 8 ⊕ 10, producing the octet baryon, has a baryon
number specified as per SU (3)c model. But not so for the fundamental entities in
the adjoint representation of baryons.
Thus it was a fundamental problem as to where does the baryon number reside
in the adjoint representation spin-1/2 baryon of the eightfold way model? We know
that in the SU (3)F model, B = 1/3 for all the three quarks (u,d,s) and hence there
is a baryon number B = 1 for the (uud) protons, for example. But still there is
no corresponding baryon number in the dual adjoint structure in the eightfold way
model.
We see here that the 0− meson octet in the Skyrme model provides the missing
baryon number to its spin-1/2 baryon partner in the eightfold way model. Thus
the Skyrme model, built up from the 0− meson octet as basic fields, topologically
produces a baryon number to play a complementary role to the spin-1/2 baryons in
the eightfold way model. Note that the eightfold way model structure of spin-1/2
baryons is global in nature and so clearly is the structure of the Skyrme model. Thus
these two complementary models - the SU (3)F model and the Skyrme model, corre-
spond to a fundamentality of the global symmetry which is dual to the local gauge
symmetry manifested in the Standard Model. So far, one has assigned elementarity
Topology for Hadrons 505

and fundamentality as a theoretical requirements to the structures arising from local


gauge symmetries. But the discussion here points to an equally significant elemen-
tarily and fundamentality as a theoretical framework for global symmetry as well.
This appears as a new duality, analogous to the well-known wave particle duality in
quantum mechanics. They complement each other and so to say, one is incomplete
without the other. And it is this Manichaean duality which, as we discussed above,
is a manifestation of the global-local complementarity paradigm.
Signficantly Skyrme model predicts the existence of a very heavy scalar sigma
meson [140]. Electric charge is also consistently quantized in the Skyrme model [141].

14.6 Appendix G: Introduction to Topology


We can define topology as the study of those properties of such spaces which
are invariant under the group of all transformations which are continuous and whose
inverses are also continuous [22]. The Euclidean line, plane, 3-space are all topological
spaces. Vaguely speaking, a continuous transformation with a continuous inverse is
one that we can achieve by bending, stretching and by twisting the space without
tearing or cutting it. Topologically one cannot tell the difference between a football
and a basketball, for one can be deformed, without tearing, to appear just like the
other. Similarly a square and a circle are topologically the same [22], [21].
Topology concerns sets for which we have enough of an idea as to when two
sets are close together, to be able to define a continuous function. Two such sets or
topological spaces are structurally the same if these is a one-to-one function map-
ping one onto the other such that both the function and its inverse are continuous.
Naively this means that our space can be stretched, shrunk and deformed in any
manner, without being torn or cut, to look just like the other. Thus a large sphere is
topologically the same structure as the small sphere, and the boundary of a circle is
the same as the boundary of a square. Two spaces are called homeomorphic when
these are structurally the same in this sense. Note that the concept of homomor-
phism is to topology as the concept of isomorphism (where the sets have the same
algebraic structure) is to algebra.
The main problem of topology is to find useful, necessary and sufficient condi-
tions for two spaces to be homomorphic. A ‘nice’ space has associated with it various
kind of groups – homology groups, cohomology groups, homotopy groups and coho-
motopy groups. If two spaces are homeomorphic it can be demonstrated that the
corresponding groups are isomorphic. Thus a necessary condition that two spaces
be homeomorphic is that their groups be isomorphic.
First a definition of open sets. On a real line R, an open interval (a, b) is defined
as the set of all points x with a < x < b as given in Figure 14.2.
Note that the end points a and b are not included in an open set. Thus the open
interval (x − δ, x + δ) surrounding x is still all in (a, b). Therefore however close x is
to a, it cannot be equal to a. Note that the empty set O is defined to be an open
set. An open set on the real line R is defined as a union of 1,2,3,.. open intervals as
depicted in Figure 14.3.
506 Group Theory in Particle, Nuclear, and Hadron Physics

a x b

FIGURE 14.2: Open interval on a real number line

FIGURE 14.3: An open set on a real number line

The same idea can be easily extended to Rd , d-dimensional space.


We define a topological space as a pair (X,Θ) of a set X and a set Θ of subsets
(as open sets) of X, such that its following axioms hold:

1. Any union of open sets is open.


2. The intersection of any two sets is open.
3. Ø and X are open.
It is said that θ is the topology of the topological space (X,Θ). However, one
leaves out the word topology from generic usage to speak simply of a topological
space X only.
A metric space is a pair (X, d) consisting of a set X and a real function d:
X · X → R which is denominated the “metric” such that
1. d(x, y) ≥ o for all x, y ∈ X and d(x, y) = 0 if and only if x = y
2. d(x, y) = d(y, x) for all x, y ∈ X
3. d(x, z) ≤ d(x, y) + d(y, z) f or any x, y, z ∈ X
This is called a triangle inequality.
One may define neighbourhood in a topological space as a pair (X, U ) con-
sisting of a set X and a family U = (Ux )x∈X of subsets of X which is called neigh-
bourhood of x then
1. Each neighbourhood of x contains x, and X is a neighbourhood of each of its
parts.
Topology for Hadrons 507

radius 6
z

radius 6
y

FIGURE 14.4: An open ball

2. Then U itself is a neighbourhood of x if U ⊂ X contains a neighbourhood of


x.
3. Each neighborhood of x is a neighbourhood of x which in turn is also a neigh-
bourhood of each of its points.
4. The intersection of any two neighbourhood of x is also a neighbourhood of x.
Next we define topology of a metric space. Given a metric space (X, d) one
takes a subset U ⊂ X as open if for every x ∈ U there is an  > 0 such that the
-ball; B (x) = {y ∈ X|d(x, y) < } centered at x is still continuous in U . The set
Θ (d) of all open sets of X is called the topology of the metric space (X, d). So we
can see that for each point y such that d(x, y) <  one may visualize a small δ− ball
entirely contained in the − ball around x as shown in Figure 14.4.
The “open ball” {y|d(x, y) < } is actually open, whence in particular a subset
U ⊂ X is a neighborhood of x if and only if it contains a ball centered on x.
A continuous map: Given X and Y as topological spaces. A map f : X → Y
is termed continuous if the inverses image of open sets is always open.
Next homeomorphism: A bijective map f : X → Y is called homomorphic
when both f and f −1 are continuous. Thus when U ⊂ X is open, if and only if
f (U ) ⊂ Y is also open. It is important to realise that the homeomorphism is to
topology what isomorphism ( where the sets have the same algebraic structure ) is
to algebra (or group theory).
Path-connectedness: X is said to be path-connected if there is a continuous
map β : [0, 1] → X such that β(0) = a and β(1) = b, i.e., every two points a, b ∈ X
are connected by a path, as shown in Figure 14.5.
Thus one sees that a path-connected space is connected. Let A and B be non-
empty, open and disjoint and X = A ∪ B. Then in X there is no path far a ∈ A and
b∈B
Invariants in topology are those entities which do not change under distortion,
i.e., which are homomorphically unchanged.
Hausdorff space: A topological space is called Hausdorff if for any two different
points there exist disjoint neighbourhoods as shown in Figure 14.6.
Note that every metric space is Hausdorff. If d is a metric with d(x, y) = δ > o
508 Group Theory in Particle, Nuclear, and Hadron Physics

ft

FIGURE 14.5: Path-connected points

V
x

U'x

FIGURE 14.6: Hausdorff space


Topology for Hadrons 509

FIGURE 14.7: Square and circle topology

then the sets Vx = z|d(x, z) < 2δ } and Vy = {z|d(y, z) < 2δ } neighbourhoods are dis-
joint. This is true for R2 space and can be generalized for Rn .
The Euclidean space R2 is Hausdorff space under the “square” or “circle” topol-
ogy. Here X = R2 with as Figure 14.7.
p
d(x, y) = (x1 − y1 )2 + (x2 − y2 )2

d(x, y) = {|x1 − y1 |, |x2 − y2 |}max


Note any two points can be separated as per the Hausdorff definition. Hence
the ‘square’ and the ‘circle’ topologies are equivalent. So very different metrics may
induce the same topology.
Hence homeomorphism exhibits significant topological predictability. Invari-
ants under homeomorphic transformation provide physically significant information
about the topological spaces.
The five Platonic solids (as used by Kepler in his ‘failed’ model to explain the
motion of planets as discussed in Chapter 1) which are solids which are obtained
using flat planes (faces) which have straight edges and the edges at points or corners
called vertices.
For polyhedra in Rn consisting of a0 vertices, a1 edges, a2 two-dimensional ‘sides’,
etc. then Euler number is

χ(p) = Σn i
i=o (−1) ai (14.40)
Thus the Euler number is a topological invariant for polyhedra P . Two examples
of polyhedra are illustrated in Figure 14.8.

For the cube a0 + a1 + a2 = 8 − 12 + 6 = 2

For the tetrahedron a0 + a1 + a2 = 4 − 6 + 4 = 2


510 Group Theory in Particle, Nuclear, and Hadron Physics

Face Vertex a

Edge

c b

Regular Cube
d
Tetrahedron (Pyramid)

FIGURE 14.8: Euler number for a cube and a tetrahedron

Let us draw the tetrahedron on a sphere as in Figure 14.9.


Yet 6 edges and 4 faces persist (faces 1,2,3 and on right plus fourth is the space
outside the new figure). It is still topologically a triangle, as it is bounded by the
same 3 edges.

————————————————–
Problem G.1: (a) Find the Euler number of the isosahedral surface.
(b) Find the Euler number of the surface S 2 of the sphere in R3
(c) Find the Euler number of a torus (a doughnut-shaped structure).
————————————————–

Betti number: One can construct a network which is all in one piece, in which
every line terminates on a free vertex. Thus it has no loops or enclosures. This defines
a tree. It will have one more vertex than there are lines or edges. Call a0 = V
(vertex), a1 = E (edges), a2 = F (faces). For R3 we see that F − E + V = 2. So for
a tree V = E + 1.
We can convert a network into a tree – i.e., leaving one connected figure. This
is done by removing some lines and edges. Suppose we remove B edges to get a tree
from a network. In Figure 14.10 (a) we remove two connected lines to obtain a tree.
So B = 2 here. V=E+1, so V=1+E-B or B=1+E-V i.e. 1+8-7 = 2. This B is the
Betti number of a network and always equals the number of faces minus 1. It is
topological invariant.
Let us look at the tetrahedron which has 6 edges, 4 vertices and 4 faces. Remove
3 edges as shown in Figure 14.9 (b) to obtain a tree and thus B = 3, which is one
less than the 4 faces.
It turns out that the Betti number bi is connected with the Euler number.
Topology for Hadrons 511

a or
4
1

a
d 2

c
c 3
b

FIGURE 14.9: Tetrahedron on a sphere

(a) (b)

FIGURE 14.10: Betti number of networks


512 Group Theory in Particle, Nuclear, and Hadron Physics

χ = Σ(−1)i bi (14.41)
But two spaces can have the same Euler number and different Betti numbers.
Hence Betti numbers are a better invariant for extracting more scientific information.
We treat the Betti number and the relevant torsion subgroup in the group
theoretical language in Chapter 2.

Differentiable Manifold: The archetypical example of a differentiable manifold


is our non-quantum mechanical model of 4-space-time R4 . Firstly, it is topological
as per our description above. Secondly, on account of the property of continuity one
is able to define differentiable functions such as are necessary to define trajectories
of the electromagnetic fields.
We can view topology as introducing two frameworks for space-time. First local
topology – this is the manner in which the concept of continuity applies here,
i.e., how open spaces fit inside one another over microscopic regions. Next the way
in which open sets are made to cover the whole space is determined by the global
topology. At present it is believed that we have a reasonably good understanding of
our space-time in terms of the local structure as that of R4 . But its global topological
structure is still an open issue.
Thus one conclusion regarding the above distinction between the local and global
topologies is that these are not related to each other in any linear manner, meaning
that one cannot be described as an approximation of some kind of the other. Thus
these frameworks exist independent of each other. This leads to the following con-
clusions:

(a) Physically we know that if an entity exists globally, then it evidently should
exist locally as well. What we are saying is that the global physical existence of an
object demands that it also exist locally. This paradigm finds resonance with the
fact that a global Lie group necessarily has a local Lie algebra [118]. This leads to a
new global-local complementarity paradigm.

(b) Next, can we make an opposite statement to the above, i.e. is it that the local
existence of an entity also demands that it exists globally as well? Let us look at the
Lie algebra-Lie group connection. It appears that, given an algebra which is a Lie
algebra, a corresponding Lie group is not necessarily guaranteed [118]. Hence there
is no local-global complementary connection. In quantum groups this is very evident
[see Chapter 13]. A quantum Lie algebra exists but it does not lead to any Lie group.
Therefore there may exist a physical reality which differs between the global and
local descriptions. For example Super-deformed Bands (SD) in nuclei (see Chapter
13), which due to their intrinsic supersymmetric property, have no counterpart in
ordinary deformed bands.
Next R4 is first a topological space – i.e., it exhibits open sets with a continu-
ous mapping. Such a space is termed a manifold. Define a function f (x) for each
point X = {x0 , . . . x3 } in R4 . Due to continuous mapping we can transform to an-
other function g defined on R4 as f (X) = g(x0 , . . . x3 ). If this function possesses
unique partial derivatives with respect to each coordinate at each point, then such
a manifold is called a differentiable manifold.
Topology for Hadrons 513

14.7 Solutions of Problems

Solution 14.1:

~ 1
U = eiφ·~τ /f = (σ + i~τ · ~π )
f
1
U† = (σ − i~τ · ~π )
f
1 1
∂µ U = (∂µ σ + i~τ · (∂µ ~π )), ∂ µ U † = (∂ µ σ − i~τ · (∂ µ ~π ))
f f
1
∂µ U ∂ µ U † = (∂µ σ + i~τ · ∂µ ~π )1(∂ µ σ − i~τ · ∂ µ ~π )1
f2
1
= [(∂µ σ)(∂ µ σ)1 + (τ · ∂µ ~π )(~τ · ∂ µ ~π )1]
f2
The second term is

= (∂µ ~π )(∂ µ ~π ) + i~τ · (∂µ ~π ) ⊗ ∂µ ~π )


The second term is zero. So,
2
= [(∂µ σ)2 + (∂µ ~π )2 ]
f2
1 1 f2
(∂µ σ)2 + (∂µ ~π )2 = T r(∂µ U ∂ µ U † )
2 2 4
————————————————–

Solution G.1:

(a) a0 + a1 + a2 = 12-30+20 = 2

(b) See Figure 14.11(a) χ(S 2 ) = a0 + a1 + a2 = 0-0+2 = 2

(c) See Figure 14.11(b) a0 + a1 + a2 = 1 − 2 + 1 = 0


(i.e., it has one vertex, 2 edges, and 1 face)
————————————————–
514 Group Theory in Particle, Nuclear, and Hadron Physics

s2

V
E

E'

FIGURE 14.11: Euler number of S 2 and a torus


Bibliography

[1] H. Weyl, Symmetry. Princeton University Press, 1952.


[2] L. Tarasov, This Amazingly Symmetrical World. Mir Publishers, Moscow,
1986.
[3] E. Wilson, Islamic Designs. British Museum Publications, London, 1988.
[4] B. Tascher, M. C. Escher - The Graphic Work. Benedikt Taschen Verlag,
Berlin GmbH, Berlin, 1990.
[5] P. Dirac, “The evolution of the physicist’s picture of nature,” Scientific Amer-
ican, vol. May 1963, 1963.
[6] T. D. Lee, Particle Physics and Introduction to Field Theory. Harwood, New
York, 1981.
[7] L. H. Ryder, Elementary Particles and Symmetries. Gordon and Breach Sci-
ence Publishers, New York, 1975.
[8] G. Galileo, II Saggiatore (The Assayer). 1616.
[9] B. Russell, The Scientific Outlook. W. W. Norton, New York, 1931.
[10] J. Jeans, The Mysterious Universe. Cambridge University Press, 1931.
[11] N. Chomsky, Syntactic Structures. Mouton, The Hague, 1957.
[12] G. Trager, The Field of Linguistics. Battenberg Press, Norman, 1946.
[13] A. Abbas, Philosophy of Science - A New Perspective. Indian Institute of
Advanced Study, Shimla, India, 2005.
[14] G. H. Hardy, A Mathematician’s Apology. Cambridge University Press, 1946.
[15] H. Margolis, Patterns, Thinking and Cognition. University of Chicago Press,
1987.
[16] A. Abbas, “Mathematics as an exact and precise language of nature,” cog-
prints.org, vol. cogprints.org/4616, 2005.
[17] R. A. Varley, N. J. C. Klessinger, C. A. J. Romanowski, and M. Siegel, “Agram-
matic but numerate,” Proc. Nat. Acad, Sci., vol. 102, p. 3519, 2005.
[18] T. W. Deacon, “Brain-language coevolution,” The evolution of human lan-
guages, SFI Studies in the science of complexity, Proceedings Volume X, ed.:
Hawkins, J. A. and Gell-Mann, M., vol. 10, 1992.
[19] D. B. Chestnut, Finite Groups and Quantum Mechanics. John Wiley and
Sons, New York, 1974.
[20] G. H. Duffey, Theoretical Physics – Classical and Modern Views. Houghton
Mifflin Company, Boston, 1973.

515
516 Bibliography

[21] J. B. Fraleigh, A Course in Abstract Algebra. Addison-Wesley Publ. Comp.,


Reading, 1971.
[22] J. P. Munkers, Topology. Prentice-Hall, 2000.
[23] K. S. Lam, Topics in Contemporary Mathematical Physics. World Scientific,
Singapore, 2003.
[24] K. Jaaenich, Topology. Prentice-Hall, 1984.
[25] J. Q. Chen, J. Ping, and F. Wang, Group Representation Theory for Physicists.
World Scientific Press, Singapore, 2002.
[26] J. Mathews and R. L. Walker, Mathematical Methods of Physics. W. A. Ben-
jamin Inc, New York, 1965.
[27] A. O. Barut and R. Raczka, Theory of Group Representation and Application.
Polish Scientific Publications, Warszawa, 1980.
[28] P. Carruthers, Introduction to Unitary Symmetry. John Wiley and Sons, New
York, 1966.
[29] M. Robinson, Symmetry and the Standard Model. Springer, Heidelberg, 2011.
[30] N. W. Dean, Introduction to the Strong Interaction. Gordon and Breach Sci-
ence Publishers, New York, 1976.
[31] R. E. Marshak, Conceptual Foundations of Modern Particle Physics. World
Scientific, Singapore, 1993.
[32] W. Greiner and B. Mueller, Quantum Mechanics - Symmetries. Springer Ver-
lag, Berlin, 1989.
[33] D. Lichtenberg, Unitary Symmetry and Elementary Particles. Academic Press,
1978.
[34] A. Abbas, “Anomalies and charge quantization in the Standard Model with
arbitraray number of colours,” Phys. Lett. B, vol. 238, p. 344, 1990.
[35] A. Abbas, “Spontaneous symmetry breaking, quantization of the electric
charge and the anomalies,” J. Phys. G, vol. 16, p. L163, 1990.
[36] A. Abbas, “Structure of A=6 nuclei: 6 He,6 Li,6 Be,” Mod. Phys. Lett. A, vol. 1,
p. 2365, 2004.
[37] A. Abbas, “What leptons really are!,” inspirehep.net/record/194660, vol. Inst.
F. Kernphysik, T.H. Darmstadt paper. IKDA83/29 (Dec. 1983), 1983.
[38] R. Laughlin, “Anomalous quantum Hall effect: an incompressible quantum
fluid with fractional charge excitations,” Phys. Rev. Lett., vol. 50, p. 1359.
[39] B. Schwarzschild, “Fractional quantum Hall effect,” Phys. Today, vol. Dec.,
p. 17, 1993.
[40] C. Cohen-Tannoudji, B. Diu, and F. Laloe, Quantum Mechanics. Wiley Inter-
sciencd, New York, 1996.
[41] W. K. Tung, Group Theory in Physics. World Scientific Press, Singapore,
1985.
[42] F. E. Close, An Introduction to Quarks and Partons. Academic Press, New
York, 1979.
Bibliography 517

[43] J. Q. Chen, “Coefficients of fractional parentage and isoscalar factors,” J.


Math. Phys., vol. 22, p. 1, 1981.
[44] D. J. Griffiths, Introduction to Quantum Mechanics. Pearson-Prentice Hall,
2005.
[45] R. Mann, An Introduction to Particle Physics and the Standard Model. CRC
Press, Boca Raton, Florida, 2010.
[46] L. H. Ryder, Quantum Field Theory. Cambridge University Press, 1985.
[47] A. Zee, Quantum Field Theory in a Nutshell. Princeton University Press, 2003.
[48] R. P. Feynman, Gauge Theories – in Weak and Electromagnetic Intercations
at High Energies, ed. Bailin and Llewellyn Smith. North Holland, Amsterdam,
1977.
[49] C. Quigg, Gauge Theories in High Energy Physics. NATO Advanced Study,
St. Croix, 1980.
[50] A. Abbas, “A new hidden colour hypothesis,” Pramana, vol. 66, p. 827, 2006.
[51] LHCb Collaboration, “Observation of j/(psi-p) resonances...,” Phys. Rev.
Lett., vol. 115, p. 072001, 2015.
[52] V. A. Matveev and P. Sorba, “Quark analysis of multibaryonic systems,” Il
Nuovo Cim. A, vol. 45, p. 257, 1978.
[53] J. Beringer and et al., “Particle Data Group,” Phys. Rev. D, vol. 86, 2012.
[54] G. S. Adkins, C. R. Nappi, and E. Witten, “Static properties of nucleon in
the Skyrme model,” Nucl. Phys. B, vol. 228, p. 552, 1983.
[55] A. H. Guth, The Inflationary Universe. Perseus Books, Reading, MA, 1997.
[56] EMC Collaboration, “A measurement of the spin asymmetry ...,” Phys. Lett.
B, vol. 206, p. 364, 1988.
[57] E142 Collaboration, “Measurement of the proton and deuteron spin
quadrupole transition...,” Phys. Rev. D, vol. 54, p. 6620, 1996.
[58] D. Flamm and F. Schoeberl, An Introduction to the Quarks and Elementary
Particles. Gordon and Breach Science Publishers, New York, 1982.
[59] S. Gasiorowicz, Elementary Particle Physics. John Wiley and Sons, 1966.
[60] S. Z. Jaunpuri, Iranianate Linguistic Influence on India Through the Ages,
vol. openlibrary.org/works/OL16118819W. Institute of Persian Research,
AMU, Aligarh, India, 2009.
[61] M. Gell-Mann and Y. Ne’eman, The Eightfold Way. W. A. Benjamin, 1964.
[62] R. N. Cahn, Semi-simple Lie Algebras and their Representations. Benjamin
Cummmins, Reading, MA, 1984.
[63] J. J. J. Kokkedee, The Quark Model. Benjamin, New York, 1968.
[64] A. N. Mitra and R. Majumdar, “Quark statistics and baryon form factors,”
Phys. Rev., vol. 150, p. 1194, 1966.
[65] M. Guidry, Gauge Field Theories – an Introduction with Applications. John
Wiley and Sons, 1999.
[66] A. Abbas, “The MIT bag model and the spin structure of the nucleons,” J.
Phys. G, vol. 15, p. L129, 1989.
518 Bibliography

[67] HERMES Collaboration, “Measurement of the neutron structure functions...,”


Phys. Lett. B, vol. 404, p. 383, 1997.
[68] E155 Collaboration, “Measurement of the Q2 -dependence of the proton ...,”
Phys. Lett. B, vol. 493, p. 19, 2000.
[69] L. M. Sehgal, “Angular momentum composition of the proton in the quark
model,” Phys. Rev. D, vol. 10, p. 1663, 1974.
[70] N. Isgur, G. Karl, and R. Koniuk, “A test of colour magnetism,” Phys. Rev.
D, vol. 25, p. 2394, 1982.
[71] S. S. Gershtein and G. V. Dzhikiya, “Quadrupole transition delta → N-gamma
in the quark model,” Sov. J. Nucl. Phys., vol. 34, p. 870, 1981.
[72] A. Abbas, “Spin of the nucleon and relativistic quarks in confining potentials,”
J. Phys. G, vol. 16, p. L21, 1990.
[73] J. S. Bell and H. Ruegg, “Dirac equation with an exact higher symmetry,”
Nucl. Phys. B, vol. 98, p. 151, 1975.
[74] R. Tegen, R. Brockmann, and W. Weise, “Chiral symmetry,” Z. Phys. A,
vol. 307, p. 339, 1982.
[75] A. Abbas, “Semi-leptonic decays of baryons in the quark model,” Europhys.
Lett., vol. 5, p. 287, 1988.
[76] A. Abbas, “Comment on colour magnetism and helicity zero γn → pδ transi-
tion amplitude,” Phys. Rev. Lett., vol. 63, p. 334, 1989.
[77] A. Abbas, “Deformed nucleon and double-delta coupling,” J. Phys. G, vol. 17,
p. L181, 1991.
[78] A. J. Buchmann and S. A. Moszkowski, “Pion couplings of the Delta (1232),”
Phys. Rev. C, vol. 87, p. 028203, 2013.
[79] A. Abbas, “Spin structure of the nucleon and the constituent quark model,”
J. Phys. G, vol. 15, p. L73, 1989.
[80] A. Abbas, “Spin content of the nucleon and the quark model,” J. Phys. G,
vol. 15, p. L195, 1989.
[81] S. A. Abbas, “Colour confinement and deformed baryons in quantum chromo-
dynamics,” J. Phys.: Conf. Series, vol. 374, p. 012006, 2006.
[82] S. Haywood, Symmetries and Conservation Laws in Particle Physics. World
Scientific, Singapore, 2011.
[83] R. E. Marshak, Riazuddin, and C. P. Ryan, Theory of Weak Interaction in
Particle Physics. Wiley Interscience, New York, 1969.
[84] F. Englert and R. Brout, “Broken symmetry and the mass of the gauge vector
mesons,” Phys. Rev. Lett., vol. 13(9), p. 321, 1964.
[85] P. Higgs, “Broken symmetries and masses of gauge bosons,” Phys. Rev. Lett.,
vol. 13(16), p. 508, 1964.
[86] G. Guralnik, C. R. Hagen, and T. W. B. Kibble, “Global conservation laws
and massless particles,” Phys. Rev. Lett., vol. 13(10), p. 555, 1964.
[87] J. Ellis, M. K. Gaillard, and D. V. Nanopoulos, “A historical profile of the
Higgs boson,” arxiv.org, vol. arxiv:1201.6045 (hep-ph), 2012.
Bibliography 519

[88] R. Brout and F. Englert, “Spontaneous symmetry breaking in gauge theories:


a historical survey,” arxiv.org, vol. arxiv:hep-th/9802142, 1998.
[89] A. Abbas, “On the Standard Model Higgs and superstring theories,” PAS-
COS99 - Particles, Strings and Cosmology, Ed. K. Cheung, J. F. Gunion and
S. Mrenna, Proc. Int. Conf., Lake Tahoe, Dec. 1999, p. 123, 2000.
[90] A. Abbas, “Standard model of particle physics has charge quantization,”
Physics Today, vol. July, p. 811, 1999.
[91] M. Maggiore, A Modern Introduction to Quantum Field Theory. Oxford Uni-
versity Press, 2005.
[92] OPERA Collaboration, “Measurement of the neutrino velocity with the
OPERA detector in the CNGS beam,” J. High Energy Phys., vol. 10, p. 093,
2012.
[93] ICARUS Collaboration, “Precision measurement of neutrino velocity...,” J.
High Energy Phys., vol. 11, p. 049, 2012.
[94] E. Giannetto, G. D. Maccarrone, R. Mignani, and E. Recami, “Are neutrinos
faster than light,” Phys. Lett. B, vol. 17, p. 115, 1986.
[95] EXO Collaboration, “Search for neutrinoless double-beta decay...,” Phys. Lett.
B, vol. 109, p. 032505, 2012.
[96] A. Abbas, “Evidence for compact dark matter in galactic halos,” PASCOS99
- Particles, Strings and Cosmology, Ed. K. Cheung, J. F. Gunion and S.
Mrenna, Proc. Int. Conf., Lake Tahoe, Dec. 1999, p. 330, 2000.
[97] S. Abbas and A. Abbas, “Volcanogenic Dark Matter and mass extinctions,”
Astropart. Phys., vol. 8, p. 317, 1998.
[98] S. S. M. Wong, Introductory Nuclear Physics. Wiley, 2013.
[99] D. M. Brink, Nuclear Forces. Oxford University Press, 1965.
[100] J. M. Blatt and V. F. Weisskopf, Theoretical Nuclear Physics. Springer Verlag,
1979.
[101] R. D. Lawson, Theory of Nuclear Shell Model. Clarendon Press, Oxford, 1980.
[102] A. Abbas and L. Zamick, “Effective charge in the large A limit,” Phys. Rev.
C, vol. 21, p. 738, 1980.
[103] A. Abbas, F. Beck, and A. Richter, “Tensor correlations and spin-isospin ex-
citations,” Z. Phys. A, vol. 321, p. 329, 1985.
[104] S. A. Abbas and S. Ahmad, “A=3 clustering in nuclei,” Int. J. Mod. Phys. E,
vol. 20, p. 2101, 2011.
[105] A. Abbas, “New proton and neutron magic numbers in nuclei,” Mod. Phys.
Lett. A, vol. 20, p. 2553, 2005.
[106] N. D. Cook, Models of the Atomic Nucleus. Springer Verlag, Berlin, 2006.
[107] A. Abbas, “Quarks and neutron halo nuclei, nuclear clusters and nuclear
molecules,” Mod. Phys. Lett. A, vol. 16, p. 755, 2001.
[108] A. Abbas, “Quasi-shell closure for even Z=58 to 70, N=82 and similarity of
these nuclei,” Phys. Rev. C, vol. 29, p. 1033, 1984.
[109] S. A. Abbas and F. H. Bhat, “Triton clustering in neutron rich nuclei,”
arxiv.org, vol. arxiv:0902.0299, 2009.
520 Bibliography

[110] P. P. Kulish and N. Y. Reshetikhin, “Quantum linear problem for the linear
sine-gordon equation and higher representations,” Zapiski Semenarov LOMI,
vol. 101, p. 101, 1981.
[111] E. K. Sklyanin, “Some algebraic structures connected with the Yang-Baxter
equations,” Funct. Anal. Appl., vol. 16, p. 262, 1982.
[112] L. C. Biedenharn, “Quantum Groups,” Lecture Notes in Physics, vol. 37, p. 67,
1990.
[113] J. Y. Zeng et al., “Spin determination and quantized spin alignment in su-
perdeformed bands ...,” Phys. Rev. C, vol. 44, p. R1745, 1991.
[114] D. Bonatsos et al., “Description of the superdeformed bands by quantum al-
gebra SU (2)q ,” J. Phys. G, vol. 17, p. L67, 1991.
[115] C. S. Wu et al., “Relation between kinematic and dynamic moments of inertia
in superdformed bands,” Phys. Rev. C, vol. 45, p. 2507, 1992.
[116] P. J. Twin, “Observation of discrete line superdeformed bands...,” Phys. Rev.
Lett., vol. 57, p. 911, 1986.
[117] J. F. Sharpey-Schaper, “Superdeformed bands in nuclei,” Prog. Part. Nucl.
Phys., vol. 28, p. 187, 1992.
[118] R. Mirman, Group Theory: An Intuitive Approach. World Scientific, Singapore,
1995.
[119] EMC Collaboration, “The ratio of the nuclear structure functions ..,”
Phys.Rev. Lett., vol. 123, p. 275, 1983.
[120] A. Abbas, “Hidden colour and the hole in the center of 3-He and 4-He,” Phys.
Lett. B, vol. 167, p. 150, 1986.
[121] R. G. Arnold et al., “Elastic electron scatteruing from ..,” Phys. Rev. Lett.,
vol. 40, p. 1429, 1978.
[122] A. Abbas, “Deformed nucleonic bags and nuclear magnetic properties,” Phys.
Rev. C, vol. 36, p. R1663, 1087.
[123] C. Gaarde, “Gamow-Teller and M1 resonances,” Nucl. Phys. A, vol. 396,
p. 127c, 1983.
[124] A. Abbas, “Gamow-Teller strengths in nuclei with deformed bags of quarks,”
J. Phys. G, vol. 13, p. 1337, 1987.
[125] N. K. Glendenning, Compact Stars. Springer Verlag, New York, 1997.
[126] B. Mueller, The Physics of Quark Gluon Plasma - Lecture Notes in Physics.
Springer Verlag, Berlin, 1985.
[127] A. Abbas, L. Paria, and S. Abbas, “On colour non-singlet representations of
the quark gluon system at finite temperature,” Eur. Phys. J. C., vol. 14, p. 675,
2000.
[128] A. Abbas and L. Paria, “Hybrids as a signature of quark gluon plasma,” J.
Phys. G., vol. 23, p. 791, 1997.
[129] L. Paria, A. Abbas, and M. G. Mustafa, “Surface tension at finite temperature
in the MIT bag model,” Int. J. Mod. Phys. E, vol. 9, p. 149, 2000.
[130] W. A. Zajc, “The fluid nature of quark gluon plasma,” Nucl. Phys. A, vol. 805,
p. 283c, 2008.
Bibliography 521

[131] U. Heinz, “Quark gluon plasma,” Int. J. Mod. Phys. A, vol. 30, p. 1530011,
2015.
[132] H. Song and U. Heinz, “Quark gluon plasma - fluid,” Phys. Lett. B, vol. 658,
p. 279, 2008.
[133] PHENIX Collaboration, “Scaling properties of azimuthal anisotropy in Au +
Au...,” Phys. Rev. Lett., vol. 98, p. 162301, 2007.
[134] K. Redlich and L. Turko, “Phase transitions in the statistical bootstrap model
with an internal symmetry,” Z. Phys. C, vol. 5, p. 201, 1980.
[135] L. Turko, “Quantum gases with internal symmetry,” Phys. Lett. B, vol. 104,
p. 153, 1981.
[136] M. I. Gorenstein, S. I. Lipshikh, V. K. Petrov, and G. M. Zinovjev, “The
colourless partition function of the quantum quark gluon gas,” Phys. Lett. B,
vol. 123, p. 437, 1983.
[137] H. T. Elze, W. Greiner, and J. Rafelski, “On the colour singlet quark gluon
plasma,” Phys. Lett. B, vol. 124, p. 515, 1983.
[138] S. A. Abbas, “An alternative framework of geometry and topology in relativity;
arxiv:0903.5532; Physical interpretation of relativity theory, (Int. Conf.) eds.
Duffy et al., BMSTU Moscow, ISBN 978-5-7038-3394-0,” 2009.
[139] R. Rajaraman, Solitons and Instantons. North Holland, Amsterdam, 1982.
[140] S. A. Abbas, “What has been discovered at 125 GeV by the CMS and the
ATLAS experiments,” Eur. Phys. J. A, vol. 49, p. 72, 2013.
[141] A. Abbas, “Does Skyrme model give a consistent description of hadrons,”
Phys. Lett. B, vol. 503, p. 81, 2001.

You might also like