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

Only $11.99/month after trial. Cancel anytime.

Group Theoretical Methods in Physics
Group Theoretical Methods in Physics
Group Theoretical Methods in Physics
Ebook978 pages7 hours

Group Theoretical Methods in Physics

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Group Theoretical Methods in Physics: Proceedings of the Fifth International Colloquium provides information pertinent to the fundamental aspects of group theoretical methods in physics. This book provides a variety of topics, including nuclear collective motion, complex Riemannian geometry, quantum mechanics, and relativistic symmetry. Organized into six parts encompassing 64 chapters, this book begins with an overview of the theories of nuclear quadrupole dynamics. This text then examines the conventional approach in the determination of superstructures. Other chapters consider the Hamiltonian formalism and how it is applied to the KdV equation and to a slight variant of the KdV equation. This book discusses as well the significant differential equations of mathematical physics that are integrable Hamiltonian systems, including the equations governing self-induced transparency and the motion of particles under an inverse square potential. The final chapter deals with the decomposition of the tensor product of two irreducible representations of the symmetric group into a direct sum of irreducible representations. This book is a valuable resource for physicists.
LanguageEnglish
Release dateDec 2, 2012
ISBN9780323141529
Group Theoretical Methods in Physics

Related to Group Theoretical Methods in Physics

Related ebooks

Physics For You

View More

Related articles

Reviews for Group Theoretical Methods in Physics

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Group Theoretical Methods in Physics - Robert Shar

    Montréal.

    Part I

    Nuclei, Atoms, Solids

    Outline

    Chapter 1: CANONICAL TRANSFORMATIONS AND SPECTRUM GENERATING ALGEBRAS IN THE THEORY OF NUCLEAR COLLECTIVE MOTION

    Chapter 2: MODULATED SPACE GROUPS

    Chapter 3: PROPERTIES OF LATTICES ASSOCIATED WITH A MODULATED CRYSTAL

    Chapter 4: A HAMILTONIAN APPROACH TO THE KdV AND OTHER EQUATIONS

    Chapter 5: USE OF AN ELEMENTARY GROUP THEORETICAL METHOD IN DETERMINING THE STRUCTURE OF A BIOLOGICAL CRYSTAL FROM ITS PATTERSON FUNCTION

    Chapter 6: INVARIANTS POLYNOMIAUX DES GROUPES DE SYMETRIE MOLECULAIRE ET CRISTALLOGRAPHIQUE

    Chapter 7: METACRYSTALLOGRAPHIC GROUPS

    Chapter 8: THE SP(3,ℝ) MODEL OF NUCLEAR COLLECTIVE MOTION

    Chapter 9: A GROUP THEORETIC DESCRIPTION OF THE MAGNETIC PHASE TRANSITIONS IN THE AB2O4-TYPE SPINELS

    Chapter 10: APPLICATIONS OF CRYSTAL CLEBSCH-GORDAN COEFFICIENTS

    Chapter 11: ON THE USE OF THE SO(4,2) DYNAMICAL GROUP FOR THE STUDY OF THE GROUND STATE OF A HYDROGEN ATOM IN A HOMOGENEOUS MAGNETIC FIELD

    Chapter 12: GROUP THEORY AROUND LIGAND FIELD THEORY

    Chapter 13: FINITE REPRESENTATIONS OF THE UNITARY GROUP AND THEIR APPLICATIONS IM MANY-BODY PHYSICS

    Chapter 14: ALGEBRAIC AND GEOMETRIC METHODS OF QUANTISATION OF THE ISOTROPIC HARMONIC OSCILLATOR

    Chapter 15: THE ISING ALGEBRA

    Chapter 16: THE GROUP AND THE HYDROGEN ATOM

    CANONICAL TRANSFORMATIONS AND SPECTRUM GENERATING ALGEBRAS IN THE THEORY OF NUCLEAR COLLECTIVE MOTION

    P. Gulshani, G. Rosensteel and D.J. Rowe

    Publisher Summary

    This chapter discusses the canonical transformations and spectrum generating algebras in the theory of nuclear collective motion. The method of canonical transformations provides valuable physical insights into the interpretation of the algebraic cm (3) model and into its relationship with the phenomenological models of quadrupole collective motions. It provides the kinetic energy component of the cm(3) Hamiltonian and observables l² and L’² that can measure the extent to which a given state describes irrotational or rigid flow. It also raises some fundamental questions regarding the nature of collective motions, namely, the impossibility of pure rigid collective flow and the impossibility of expressing a many-particle wave function in terms of the collective and intrinsic coordinates. All the variables in the kinetic energy are well-defined and have a well-defined action on many-particle Hubert space. However, it does mean that a given wave function, expressed in terms of particle coordinates, cannot be reexpressed in terms of the cm, collective, and intrinsic coordinates.

    Theories of nuclear quadrupole dynamics fall essentially into two classes: phenomenological models, which are expressed in terms of ad hoc collective coordinates, and microscopic theories which attempt to explain collectivity in terms of coherent motions of particles [1]. With the algebraic models came the means of relating the two. Thus, for example, in the [R⁵]so(3) model [2], one can identify the abelian subalgebra R⁵ with the 5 components of the traceless mass quadrupole tensor and so(3) with the rotational angular momentum. In this way one obtains the phenomenological rotational model. But at the same time the algebra [R⁵] so(3) has a well-defined action on particle coordinates and so one has the beginnings of a microscopic theory.

    In addition to describing quadrupole dynamics, one would like a theory which would also predict, or at least provide the means to observe, what goes on inside a rotational nucleus. For that we need to learn what are the relevant quantum mechanical observables that could, for example, distinguish between some of the possible flow patterns illustrated in Fig. (1). We pursue these questions and the relationships between the phenomenological and microscopic collective models by making a linear canonical point transformation from particle coordinates to centre-of-mass, collective and intrinsic coordinates.

    Fig. 1 Possible nuclear flow patterns (a) irrotational flow (b) rigid flow (c) two-fluid flow.

    The method is described in detail in Ref. [3]. A canonical transformation is made in two steps.

    is the c.m. coordinate, g ε GL(3,R) is a function of 9 collective coordinates θ, λ, ψ and xnα" is a function of 3N-12 intrinsic coordinates ξ. When the corresponding canonical transformation is applied to the momentum coordinates, one obtains a separation of the Hamiltonian

    are defined in the usual manner so that there is no term in the Hamiltonian coupling the relative and c.m. degrees of freedom. We therefore define the collective coordinates by the parallel criterion of minimizing Hcoup.

    Consider the Cartan decomposition of g

    , where M is the mass of a particle, such that there is no quadrupole deformation in the intrinsic system. Thus SO is a scale transformation on each of the three principal coordinate axes. Finally R(ψ) is chosen such that there is zero angular momentum of the system relative to the intrinsic coordinate axes. As a consequence of this latter choice, we find that the Coriolis force vanishes and that the kinetic energy becomes

    with

    the angular momenta LAB which act on θ and the vorticity operators £AB, which are the angular momenta acting on π. These quantities are all in the enveloping algebra of cm(3) [4]. This is a highly significant result. For one thing it strongly supports the cm(3) model’s status as the relevant algebraic model for quadrupole collective motions. For another, it supplies the appropriate kinetic energy for the cm(3) model.

    Recall that the cm(3) algebra contains the quadrupole mass tensor, the angular momenta which generate rigid rotations, and the shear operators which generate irrotational flow vibrational and rotational displacements. Thus the cm(3) model contains the potentiality for describing rigid flow, irrotational flow and all possible linear combinations of the two. It is of interest therefore to determine what flow patterns correspond to the various irreducible representations of cm(3).

    The transformed expression for the momentum of a particle has three components

    by

    are operators. We can nevertheless extend the analysis to quantum mechanics in the natural way and say that a representation of cm(3) describes irrotational flow if

    for all states ψ in the carrier space of the irreducible representation. Since £² is one of the quadratic invariants of cm(3) [3,4,5] there is a large class of representations for which this condition holds.

    should vanish, where

    In quantum mechanics then we say that a state ψ describes rigid collective flow if

    do not by themselves form a closed algebra and thus it is not, in general, possible to satisfy the conditions for rigid flow. Thus it appears that rigid flow can occur only in exceptional circumstances or under certain limiting conditions, e.g., a classical limit, which remain to be investigated.

    This is not a problem for the canonical transformation of the kinetic energy, since the coordinates ψ are cyclic (i.e., they do not appear explicitly in the kinetic energy). Furthermore, all the variables in the kinetic energy are well-defined and have a well-defined action on many-particle Hilbert space. However, it does mean that a given wave-function, expressed in terms of particle coordinates, cannot be re-expressed in terms of the above c.m., collective and intrinsic coordinates. The full implications of this observation for the microscopic theory of collective motion remain to be investigated.

    REFERENCES

    1. ROWE, D. J.Nuclear Collective Motion; Models and Theory. London: Methuen, 1970.

    2. UI, H. Prog. Theor. Phys. 1970; 44:153. WEAVER, O. L., BIEDENHARN, L. C., CUSSON, R. Y. Ann. Phys. (New York). 1973; 77:250.

    3. GULSHANI, P., ROWE, D. J. Can. Journ. Phys. 1976; 54:970.

    4. ROSENSTEEL, G., ROWE, D. J. Ann. Phys. (New York). 1976; 96:1.

    5. O.L. WEAVER, Factorization of the Invariant Operators of CM(3), contributed paper, session 1A.

    MODULATED SPACE GROUPS

    A. Janner

    Publisher Summary

    This chapter discusses modulated space groups. The elementary supercell of a crystal with a superstructure has a finite volume. If this supercell is of macroscopic size, one gets a modulated crystal structure. A modulated crystal has an infinite elementary cell, which is considered as the limiting case of a larger and larger supercell. In the case of modulated crystals, the Euclidean symmetry of the corresponding basic pattern is that of a space group of the same dimension as the crystal. The modulation destroys this property. The remaining Euclidean symmetry does not explain systematic extinctions observed in the diffraction pattern of modulated crystals. One can make use of the space group symmetry of the basic structure and treat the modulation as a perturbation. This is the conventional approach in the determination of superstructures, which is insofar not a satisfactory one as the basic structure does not always have a physical meaning and is not uniquely determined.

    1 INTRODUCTION

    Consider a crystal with a superstructure. Its elementary supercell has a finite volume. If this supercell is of macroscopic size, one gets a modulated crystal structure. Mathematically speaking, however, a modulated crystal has an infinite elementary cell, which can be considered as the limiting case of a larger and larger supercell. Indeed by modulation we mean a periodic deviation from a basic periodic pattern, both periodicities being incommensurate. In the case of modulated crystals the Euclidean symmetry of the corresponding basic pattern is that of a space group of the same dimension as the crystal. The modulation destroys this property. Furthermore the remaining Euclidean symmetry (if any) does not explain systematic extinctions observed in the diffraction pattern of modulated crystals [see e.g. ref. 1 and 2]. One can of course make use of the space group symmetry of the basic structure and treat the modulation as a perturbation. This is the conventional approach in the determination of superstructures³ which is insofar not a satisfactory one, as the basic structure does not always have a physical meaning and, in general, is not uniquely determined.

    In an alternative approach, the modulated crystal is considered as section of a higher-dimensional periodic structure. Under certain assumptions (of continuity e.g.) the imbedding of the modulated crystal in the higher dimensional Euclidean space is essentially unique and the symmetry of the imbedding crystal (which is a higher dimensional space group) does explain systematic diffractive extinctions [see ref. 4 and 5 for more details]. Going further along this line, it will be shown here, that nevertheless, the symmetry of the modulated crystal can equally well be described by a space group of the same dimension as that of the crystal: This group, called modulated space group, however, is not a Euclidean symmetry group. The situation is very much the same as that of the point symmetry of a normal crystal, if its space group is non-symmorphic, because then the point group is not the orthogonal symmetry group of such a crystal. This comparison will be discussed in more details in the last section.

    Let us here recall some basic properties of normal and of modulated crystals fixing at the same time the notation. A general treatment makes the formal structure more transparent; we therefore consider a n-dimensional crystal (the n=3 case being the most important one from the physical point of view) described by a scalar function:

    (1)

    (k) ≠ 0. Here Sp generates Vn.

    A.

    In a normal crystal the elements of S can be written as:

    (2)

    ) for g ε E(n)) is then a n-dimensional space group⁶, which is defined by the following properties:

    (3)

    where E(n) is the n-dim. Euclidean group, Tn its subgroup of translations, and the group Λn of lattice translations (identified after choice of an origin with the lattice of points equivalent with the chosen origin) generates the whole space Vn. Using Seitz’s notation⁷, the elements of G can be written as:

    (4)

    ε Vn given by:

    a system of non-primitive translations, whose properties are:

    (5)

    (R) depends on the choice of the origin and is determined only modulo primitive translations.)

    are given. (See ref. 8 for more details.)

    B.

    In a modulated crystal the elements of Sp can be written as:

    (6)

    a basis Bd* of a d-dimensional lattice Dd*. (Confusion, say in the case n = d = 3, can be avoided by writing B*n=3 and Bd=3*.

    The periodicity of the modulation implies 1 ≤ d ≤ n, and its incommensurability with respect to Λn* can be expressed (without restriction of generality) by the condition:

    (7)

    are called main reflections: those for which this is not the case are called satellite reflections. Note that the main reflections occur at the lattice points of Λn* whereas the satellite reflections occur at the points of lattices Dd*(Λn*) generated by Dd* from each point of Λn*.

    Clearly the description of Sp in terms of the bases Bn* and Bd* as in (6) implies a choice. The validity of the essential properties derived hereafter is independent of this choice. To get this non trivial result requires a careful inspection of the consequences of this arbitrariness. A first discussion on it can be found elsewhere in these same proceedings⁹. A more complete analysis will be published elsewhere.

    By the following proposition the imbedding idea mentioned above becomes natural.

    Proposition 1.

    implies that the abelian group (Λn*,Dd*) freely generated by the bases Bn* and Bd* is isomorphic to Zn+d:

    (8)

    ) in a (n+d)-dim. space.

    2 IMBEDDING

    The lattice Λn* generates Vn (as Euclidean space) and Dd* generates the (non trivial) Euclidean d-dimensional subspace Vd ⊆ Vn.

    We now consider the (n+d)-dim. Euclidean space VS given by:

    (9)

    We call VE (isomorphic with Vn) the external space, VI (isomorphic with Vd) the internal space and Vn+d = VS the superspace.

    The metric gijE in VE is the same as that in Vn whereas the metric gijI in VI need not to be that of Vd†) and depends on the choice of the lattice Dd* among all possible Dd*(Λn*). In the cases where it is desirable to take this explicitly into account we will denote the scalar product in VE by a dot and that in VI by an open circle. Usually, however, these symbols will be omitted.

    We adopt the notation:

    and scalar product

    (10)

    (Note that in ref. 5 the minus sign was adopted because it was more convenient in that context.)

    As (standard) basis B*n+d for VS we define:

    (11)

    This ensures that the projection πE of the abelian group Σn+d*, freely generated by Bn+d*, on its external components is (Λn*,Dd*), and that VI is a copy of Vd. From given as in (6):

    (12)

    so that the function ρ describing the modulated crystal in Vn can be extended to one in VS:

    (13a)

    be defining:

    (13b)

    so that

    (14)

    3 SUPERSPACE GROUP G

    (g−1r).

    Accordingly its symmetry group G is given by:

    (15)

    We adopt the notation:

    (16)

    Then the symmetry condition expressed in terms of the Fourier components becomes:

    (17)

    where dot indicates here the scalar product in VS.

    Proposition 2.

    G is a (n+d)-dimensional space group of the superspace VS.

    Proof.

    Zn+d and Σn+d generates VS. We therefore call G a superspace group.

    Knowing that G is a space group, we analyse further its group of lattice translations, its point group K and its system of non-primitive translations v.

    reciprocal to Bn* and to Bd* in VE and V1 respectively, one finds for Bn+d, the basis reciprocal in VS to Bn+d*:

    (18)

    where the condition:

    (19)

    implies

    (20)

    Vd, the basis Bd of VI can be identified with the corresponding basis of Vd and so also Bd*. Even if Bd and Bd* are no more dual in Vd, ), and is of rank d (Tilde means transposed).

    Identifying groups of lattice translations with the corresponding lattices (after choice of a fixed origin) and denoting by Λn, Dd the lattices reciprocal to Λn* and Dd* in VE and VI, respectively, by considering (11) and (18) one immediately gets:

    (21)

    (22)

    Furthermore identifying (0,Dd) with Dd and (Λn*,0) with Λn* we also have:

    (23)

    πE and πI denoting, when acting on VS, the orthogonal projection on VE and VI respectively, and when acting on group elements of G, the corresponding projection on their external and internal components.

    Proposition 3.

    where TId denotes the group of internal translations.

    Proof.

    by proposition

    2. The result then follows from (21).

    The point group K of the space group G is defined as the group of the homogeneous parts of the elements of G:

    According to the general theory on space groups one then knows that K leaves Σn+d and Σn+d*

    Enjoying the preview?
    Page 1 of 1