Representations of the Lorentz and Poincare Groups and Transformation Properties of Physical Quantities

Total Page:16

File Type:pdf, Size:1020Kb

Representations of the Lorentz and Poincare Groups and Transformation Properties of Physical Quantities CHAPTER 2 Representations of the Lorentz and Poincare Groups and Transformation Properties of Physical Quantities At the very outset, it is pertinent to mention that finite dimensional irreducible representations (IRRs) of both these groups are not unitary and all unitary representations are infinite dimensional since the groups are non-compact. 2.1 UNITARY REPRESENTATIONS OF THE LORENTZ GROUP (WIGNER, 1939; GEL'FAND, 1963; NAIMARK, 1964; OHNUKI, 1976; TUNG, 1984) 2.1.1 Reduction of Representations and Relabelling of Basis The Lie algebra for the proper Lorentz group is given by JJμλσv , = iejJ σvvg μλ J λg μσ J μλg v σ+ J μσg v λ (1) or equivalently by mnl JJmn, = ie J l (2) mnl KJmn, = ie K l (3) mnl KKmn, = ie J l (4) Representations of the Lorentz and Poincare Groups and Transformation Properties... 49 related to the eigenvalue of the Casimir operator M 2 and write eq. (12) as M 3 |ukÒ = k|ukÒ (13) and, further 2 M |ukÒ = u(u + 1)|ukÒ (14) Of these, the index u corresponding to the eigenvalue of the Casimir operator M 2 labels out the IRR, since it is left invariant by group operations and hence, is constant across a given IRR, whereas the other index k, which is the eigenvalue of M 3 labels out the basis vectors within an IRR. We have M M |uk = M M |uk + MM , uk 3 ± Ò ± 3 Ò 3 ± = M ± M 3|ukÒ ± M ± |ukÒ = (k ± 1) M ± |ukÒ (15) Thus, M ± |ukÒ is an eigenstate of M 3 with eigenvalue k ± 1. This means that, in line with our ± labelling scheme for eigenstates, we must have M ± |ukÒ•|u, k ± 1Ò or M ± |ukÒ = Am|u, k ± 1Ò. Similar calculations apply to SU(2)N. Hence, we can label the IRRs of the direct product algebra SU(2)M ƒ SU(2)N that provides the IRRs for the Lorentz group by two numbers u, v corresponding to the two Casimir operators M 2 2 and N of the subalgebras SU(2)M and SU(2)N respectively. The basis vectors in the representation space are, then, chosen as {|uv; klÒ} ∫ {|klÒ} that consist of the direct products of the basis vectors from the canonical basis {|ukÒ} and {|vlÒ} of the two subalgebras SU(2)M and SU(2)N respectively. On this basis, the action of the generators of the Lorentz group is given by the following equations: |kl = MNkl = (k + l)|kl (16) J 3 Ò ej33+ Ò J kl = MNkl = A±(u, k)|k ± 1, l + B± (v, l)|k, l ± 1 (17) ± ej±±+ Ò Ò K 3 kl = iejNM33 kl = i(l k|klÒ (18) K kl = i NM kl = iA± (u, k)|k ± 1, l + iB± (v, l)|k, l ± 1 (19) ± ej±± Ò Ò It is seen from Eqs. (16) and (17) that the restriction of the representation space to the subgroup of three-dimensional rotations is equivalent to that of a direct product of representations of the SO(3) group labelled by the indices u and v respectively. We can reduce this direct product of representations to a direct sum of invariant subspaces with respect to rotations. For this purpose, we identify these constituents of the direct product representations in terms of u, v corresponding the eigenvalues of the respective Casimir operators and denote them by Du [R] and Dv[R] respectively. We denote the direct product of these two 50 Relativistic Quantum Mechanics representations by Du × v[R]. We designate, as above, the basis vectors in the product u × v representations by {|uv; klÒ} ∫ {|klÒ} so that |klÒ = |ukÒ ƒ |vlÒ. The action of D [R] on the direct product space W = VM ƒ VN is obtained by its action on the direct product basis |klÒ = |ukÒ ƒ |vlÒ, e.g., u × v u m v n D [R]|klÒ = D [R] k D [R] l |mnÒ (20) We show that Du × v [R] defines a representation of SO(3). This can be easily done by u + v u × v computing the character of the direct product representation. We have c = TrD [R] u × v k u i v j u v = D [R]k = D [R]i D [R] j = c c establishing our proposition. We examine the reducibility of Du × v[R]. For this purpose, we write u × v λ D [R]~ ∑ aDλ [] R (21) λ⊕ i.e., we attempt to reduce the direct product space W into a direct sum of invariant subspaces W l, a where labels out the IRR and =1, 2,... , distinguishes the spaces that correspond to the l a al same l. We can, then, choose a basis in W corresponding to any IRR, i.e., any l, such that the l l l l first n1 basis vectors are in W 1, the next n1 in W 2, and so on. Therefore, the representation matrices are in block diagonal form in the new basis. To examine the reducibility of our direct product representation, we consider an infinitesimal rotation about an arbitrary axis n . Then, we must have u v u × v D [Rn(dq)] D [Rn(dq)] = D [Rn(dq)] (22) whence, to first order in dq u u v v u v uvu v uv×× uv ejI iJdn θ . ejI iJdn θ = I ƒ I iejJJn ⊗+⊗II n dq = ejI iJdn θ (23) uv× u v u v which gives J n = J n ƒ I ¥ I ƒ J n . Thus, we find that the generators of a direct product representation are the sums of the corresponding generators of its constituent representations. uv× u v u v The generators defined by J n = J n ƒ I + I ƒ J n satisfy the same Lie algebra as the uv×× uv m uv¥ original one, i.e., JJk , l = iekl Jm . We have uv×× uv uvu v uvu v JJk , l = ejejJJJJk ƒ+ƒIIk , l ƒ+ƒ II l = JJuvu⊗+⊗II v JJuvu⊗+⊗II v ejk k ejl l uvu v uvu v ejJJl ⊗+⊗II l ejJJk ⊗+⊗II k uvuv u v u v = ejJ k ⊗ I ejJ l ⊗ I + ejI ⊗ J k ejI ⊗ J l uvuv u v u v ejJ l ⊗ I ejJ k ⊗ I ejI ⊗ J l ejI ⊗ J k Representations of the Lorentz and Poincare Groups and Transformation Properties... 51 u u v u v v u u v u v v = ejJJk l ƒ I + I ƒ ejJJk , l ejJJl , k ƒ I I ƒ ejJJl , k u u v u v v m uvu v m uv× = JJk , l ƒ I + I ƒ JJk , l = iekl ejJJm ⊗+⊗II m = iekl J m Our next step is to regroup the basis vectors of the product representation {|klÒ} to form invariant subspaces thereby leading to the reduction of the direct product representation. We also have uv¥ uvu v J3 kl = ejJJ33×+II × (|ukÒ × |vlÒ) = (k + l) (|ukÒ × |vlÒ) = (k + l)|klÒ (24) uv× so that |klÒ is an eigenvector of J 3 with eigenvalue (k + l). 2 uv× 2 uv× uv× uv× uv× Further, (J ) = ejJ 3 + J 3 + J J + whence 2 u × v uvuvu v u v ejJ uk ƒ|vl Ò = ejJJ33⊗+⊗II ejJJ33ƒ+ƒII uk ƒ|vlÒ + JJu ⊗+⊗IIv u v uk |vl + JJuvuv⊗+⊗II JJuvuvII uk |vl ej33ƒ Ò ejej++ƒ+ƒ ƒ Ò 2 uvuv = (k + l) |ukÒ ƒ|vlÒ + (k + l)|ukÒ ƒ |vlÒ + ejJJ⊗+⊗II + + [A (u, k)|u, k + 1Ò ƒ|v, lÒ + B (v, l)|uk ƒ|v, l + 1Ò] 2 = (k + l) |ukÒ ƒ|vlÒ + (k + l)|ukÒ ƒ |vlÒ + A (u, k + 1) + + A (u, k)|u, kÒ ƒ|v, lÒ + B (v, l + 1)B (v, l)|u, kÒ ƒ|v, lÒ 2 so that |ukÒ ƒ |vlÒ is also an eigenvector of ejJ uv× . We also know from the elementary theory of angular momentum in three dimensions, of which the group SO(3) provides a representation, that the range of the eigenvalues of the generator u 2 M ∫ J 3 extends from u £ k £ + u, where u relates to the eigenvalue of the Casimir operator M . uv× Similarly, the range of l extends from v £ l £ + v. Therefore, the highest eigenvalue of J 3 is u + v and there is only one eigenvector corresponding to this eigenvalue given by |uvÒ. There will be uv× two eigenvectors corresponding to the eigenvalue u + v 1 of J 3 given by |u 1, vÒ and |u, v 1Ò uv¥ and so on. The lowest eigenvalue of J3 is u v and there is only one eigenvector corresponding uv¥ to this eigenvalue given by | u, vÒ. Since the eigenvalues of J3 change in integral steps, the total uv¥ number of eigenvalues of J3 will be 2(u + v) + 1 and the corresponding eigenvectors can be uv× obtained by applying the lowering operator J to the highest state |uvÒ. 52 Relativistic Quantum Mechanics Since there is only one state with the eigenvalue u + v and there exists no state with a higher u ¥ v eigenvalue of J 3 , this state i.e. |uvÒ is the highest member of the basis of the IRR labelled by u + v. The remaining basis vectors can be generated by successive application of the lowering operator u¥v J , as mentioned above. The total number of basis vectors in this representation will be u ¥ v 2(u + v) + 1 (because the range of eigenvalues of J 3 is from u + v through 0 to (u + v)) and they constitute the basis of an invariant subspace corresponding to the IRR labelled by u + v relating it to the eigenvalue of ( J u ¥ v)2.
Recommended publications
  • The Unitary Representations of the Poincaré Group in Any Spacetime
    The unitary representations of the Poincar´e group in any spacetime dimension Xavier Bekaert a and Nicolas Boulanger b a Institut Denis Poisson, Unit´emixte de Recherche 7013, Universit´ede Tours, Universit´ed’Orl´eans, CNRS, Parc de Grandmont, 37200 Tours (France) [email protected] b Service de Physique de l’Univers, Champs et Gravitation Universit´ede Mons – UMONS, Place du Parc 20, 7000 Mons (Belgium) [email protected] An extensive group-theoretical treatment of linear relativistic field equa- tions on Minkowski spacetime of arbitrary dimension D > 2 is presented in these lecture notes. To start with, the one-to-one correspondence be- tween linear relativistic field equations and unitary representations of the isometry group is reviewed. In turn, the method of induced representa- tions reduces the problem of classifying the representations of the Poincar´e group ISO(D 1, 1) to the classification of the representations of the sta- − bility subgroups only. Therefore, an exhaustive treatment of the two most important classes of unitary irreducible representations, corresponding to massive and massless particles (the latter class decomposing in turn into the “helicity” and the “infinite-spin” representations) may be performed via the well-known representation theory of the orthogonal groups O(n) (with D 4 <n<D ). Finally, covariant field equations are given for each − unitary irreducible representation of the Poincar´egroup with non-negative arXiv:hep-th/0611263v2 13 Jun 2021 mass-squared. Tachyonic representations are also examined. All these steps are covered in many details and with examples. The present notes also include a self-contained review of the representation theory of the general linear and (in)homogeneous orthogonal groups in terms of Young diagrams.
    [Show full text]
  • Group Theory
    Appendix A Group Theory This appendix is a survey of only those topics in group theory that are needed to understand the composition of symmetry transformations and its consequences for fundamental physics. It is intended to be self-contained and covers those topics that are needed to follow the main text. Although in the end this appendix became quite long, a thorough understanding of group theory is possible only by consulting the appropriate literature in addition to this appendix. In order that this book not become too lengthy, proofs of theorems were largely omitted; again I refer to other monographs. From its very title, the book by H. Georgi [211] is the most appropriate if particle physics is the primary focus of interest. The book by G. Costa and G. Fogli [102] is written in the same spirit. Both books also cover the necessary group theory for grand unification ideas. A very comprehensive but also rather dense treatment is given by [428]. Still a classic is [254]; it contains more about the treatment of dynamical symmetries in quantum mechanics. A.1 Basics A.1.1 Definitions: Algebraic Structures From the structureless notion of a set, one can successively generate more and more algebraic structures. Those that play a prominent role in physics are defined in the following. Group A group G is a set with elements gi and an operation ◦ (called group multiplication) with the properties that (i) the operation is closed: gi ◦ g j ∈ G, (ii) a neutral element g0 ∈ G exists such that gi ◦ g0 = g0 ◦ gi = gi , (iii) for every gi exists an −1 ∈ ◦ −1 = = −1 ◦ inverse element gi G such that gi gi g0 gi gi , (iv) the operation is associative: gi ◦ (g j ◦ gk) = (gi ◦ g j ) ◦ gk.
    [Show full text]
  • Matrix Lie Groups
    Maths Seminar 2007 MATRIX LIE GROUPS Claudiu C Remsing Dept of Mathematics (Pure and Applied) Rhodes University Grahamstown 6140 26 September 2007 RhodesUniv CCR 0 Maths Seminar 2007 TALK OUTLINE 1. What is a matrix Lie group ? 2. Matrices revisited. 3. Examples of matrix Lie groups. 4. Matrix Lie algebras. 5. A glimpse at elementary Lie theory. 6. Life beyond elementary Lie theory. RhodesUniv CCR 1 Maths Seminar 2007 1. What is a matrix Lie group ? Matrix Lie groups are groups of invertible • matrices that have desirable geometric features. So matrix Lie groups are simultaneously algebraic and geometric objects. Matrix Lie groups naturally arise in • – geometry (classical, algebraic, differential) – complex analyis – differential equations – Fourier analysis – algebra (group theory, ring theory) – number theory – combinatorics. RhodesUniv CCR 2 Maths Seminar 2007 Matrix Lie groups are encountered in many • applications in – physics (geometric mechanics, quantum con- trol) – engineering (motion control, robotics) – computational chemistry (molecular mo- tion) – computer science (computer animation, computer vision, quantum computation). “It turns out that matrix [Lie] groups • pop up in virtually any investigation of objects with symmetries, such as molecules in chemistry, particles in physics, and projective spaces in geometry”. (K. Tapp, 2005) RhodesUniv CCR 3 Maths Seminar 2007 EXAMPLE 1 : The Euclidean group E (2). • E (2) = F : R2 R2 F is an isometry . → | n o The vector space R2 is equipped with the standard Euclidean structure (the “dot product”) x y = x y + x y (x, y R2), • 1 1 2 2 ∈ hence with the Euclidean distance d (x, y) = (y x) (y x) (x, y R2).
    [Show full text]
  • LECTURE 12: LIE GROUPS and THEIR LIE ALGEBRAS 1. Lie
    LECTURE 12: LIE GROUPS AND THEIR LIE ALGEBRAS 1. Lie groups Definition 1.1. A Lie group G is a smooth manifold equipped with a group structure so that the group multiplication µ : G × G ! G; (g1; g2) 7! g1 · g2 is a smooth map. Example. Here are some basic examples: • Rn, considered as a group under addition. • R∗ = R − f0g, considered as a group under multiplication. • S1, Considered as a group under multiplication. • Linear Lie groups GL(n; R), SL(n; R), O(n) etc. • If M and N are Lie groups, so is their product M × N. Remarks. (1) (Hilbert's 5th problem, [Gleason and Montgomery-Zippin, 1950's]) Any topological group whose underlying space is a topological manifold is a Lie group. (2) Not every smooth manifold admits a Lie group structure. For example, the only spheres that admit a Lie group structure are S0, S1 and S3; among all the compact 2 dimensional surfaces the only one that admits a Lie group structure is T 2 = S1 × S1. (3) Here are two simple topological constraints for a manifold to be a Lie group: • If G is a Lie group, then TG is a trivial bundle. n { Proof: We identify TeG = R . The vector bundle isomorphism is given by φ : G × TeG ! T G; φ(x; ξ) = (x; dLx(ξ)) • If G is a Lie group, then π1(G) is an abelian group. { Proof: Suppose α1, α2 2 π1(G). Define α : [0; 1] × [0; 1] ! G by α(t1; t2) = α1(t1) · α2(t2). Then along the bottom edge followed by the right edge we have the composition α1 ◦ α2, where ◦ is the product of loops in the fundamental group, while along the left edge followed by the top edge we get α2 ◦ α1.
    [Show full text]
  • Representations of the Euclidean Group and Its Applications to the Kinematics of Spatial Chains
    REPRESENTATIONS OF THE EUCLIDEAN GROUP AND ITS APPLICATIONS TO THE KINEMATICS OF SPATIAL CHAINS By JOSE MARIA RICO MARTINEZ A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY UNIVERSITY OF FLORIDA 1988 LIBRARIES v ||miyBRSlTY_OF FLORIDA Copyright 1988 by Jose Maria Rico Martinez ACKNOWLEDGMENTS The author wishes to thank firstly his advisor Dr. Joseph Duffy for his guidance during the selection and development of the contents of this dissertation. Without his encouragement, in times when nothing fruitful seemed to evolve from the approaches followed, without his geometrical insight, and his quest for clarity, this dissertation would have plenty of errors. Nonetheless, the author is to blame for the remaining ones. Secondly, deep felt thanks go to the members of the supervisory committee for their invaluable criticism and for their teachings in the classroom. The faculty of the Mechanical Engineering Department, and in particular the faculty of the Center for Intelligent Machines and Robotics, must be thanked for the development of an atmosphere conductive to research. The contributions of the faculty of this center, including the visiting professors, can be found in many parts of this work. Special gratitude is owed to visiting professors Eric Primrose and Kenneth H. Hunt for their insightful comments. My thanks go also to my fellow students for their friendship and kindness. The economic support from the Mexican Ministry of Public Education and the Consejo Nacional de Ciencia y Tecnologia (CONACYT) is dutifully acknowledged. Finally, the author thanks his wife iii .
    [Show full text]
  • 1 the Spin Homomorphism SL2(C) → SO1,3(R) a Summary from Multiple Sources Written by Y
    1 The spin homomorphism SL2(C) ! SO1;3(R) A summary from multiple sources written by Y. Feng and Katherine E. Stange Abstract We will discuss the spin homomorphism SL2(C) ! SO1;3(R) in three manners. Firstly we interpret SL2(C) as acting on the Minkowski 1;3 spacetime R ; secondly by viewing the quadratic form as a twisted 1 1 P × P ; and finally using Clifford groups. 1.1 Introduction The spin homomorphism SL2(C) ! SO1;3(R) is a homomorphism of classical matrix Lie groups. The lefthand group con- sists of 2 × 2 complex matrices with determinant 1. The righthand group consists of 4 × 4 real matrices with determinant 1 which preserve some fixed real quadratic form Q of signature (1; 3). This map is alternately called the spinor map and variations. The image of this map is the identity component + of SO1;3(R), denoted SO1;3(R). The kernel is {±Ig. Therefore, we obtain an isomorphism + PSL2(C) = SL2(C)= ± I ' SO1;3(R): This is one of a family of isomorphisms of Lie groups called exceptional iso- morphisms. In Section 1.3, we give the spin homomorphism explicitly, al- though these formulae are unenlightening by themselves. In Section 1.4 we describe O1;3(R) in greater detail as the group of Lorentz transformations. This document describes this homomorphism from three distinct per- spectives. The first is very concrete, and constructs, using the language of Minkowski space, Lorentz transformations and Hermitian matrices, an ex- 4 plicit action of SL2(C) on R preserving Q (Section 1.5).
    [Show full text]
  • Physics of the Lorentz Group
    Physics of the Lorentz Group Sibel Ba¸skal Department of Physics, Middle East Technical University, 06800 Ankara, Turkey e-mail: [email protected] Young S. Kim Center for Fundamental Physics, University of Maryland, College Park, Maryland 20742, U.S.A. e-mail: [email protected] Marilyn E. Noz Department of Radiology, New York University, New York, NY 10016 U.S.A. e-mail: [email protected] To be published in the IOP Concise Book Series. Preface When Newton formulated his law of gravity, he wrote down his formula applicable to two point particles. It took him 20 years to prove that his formula works also for extended objects such as the sun and earth. When Einstein formulated his special relativity in 1905, he worked out the transformation law for point particles. The question is what happens when those particles have space-time extensions. The hydrogen atom is a case in point. The hydrogen atom is small enough to be regarded as a particle obeyingp Einstein's law of Lorentz transformations including the energy-momentum relation E = p2 + m2. Yet, it is known to have a rich internal space-time structure, rich enough to provide the foundation of quantum mechanics. Indeed, Niels Bohr was interested in why the energy levels of the hydrogen atom are discrete. His interest led to the replacement of the orbit by a standing wave. Before and after 1927, Einstein and Bohr met occasionally to discuss physics. It is possible that they discussed how the hydrogen atom with an electron orbit or as a standing-wave looks to moving observers.
    [Show full text]
  • Physics of the Lorentz Group
    Physics of the Lorentz Group Physics of the Lorentz Group Sibel Ba¸skal Department of Physics, Middle East Technical University, 06800 Ankara, Turkey e-mail: [email protected] Young S. Kim Center for Fundamental Physics, University of Maryland, College Park, Maryland 20742, U.S.A. e-mail: [email protected] Marilyn E. Noz Department of Radiology, New York University, New York, NY 10016 U.S.A. e-mail: [email protected] Preface When Newton formulated his law of gravity, he wrote down his formula applicable to two point particles. It took him 20 years to prove that his formula works also for extended objects such as the sun and earth. When Einstein formulated his special relativity in 1905, he worked out the transfor- mation law for point particles. The question is what happens when those particles have space-time extensions. The hydrogen atom is a case in point. The hydrogen atom is small enough to be regarded as a particle obeying Einstein'sp law of Lorentz transformations in- cluding the energy-momentum relation E = p2 + m2. Yet, it is known to have a rich internal space-time structure, rich enough to provide the foundation of quantum mechanics. Indeed, Niels Bohr was interested in why the energy levels of the hydrogen atom are discrete. His interest led to the replacement of the orbit by a standing wave. Before and after 1927, Einstein and Bohr met occasionally to discuss physics. It is possible that they discussed how the hydrogen atom with an electron orbit or as a standing-wave looks to moving observers.
    [Show full text]
  • Quantum Mechanics 2 Tutorial 9: the Lorentz Group
    Quantum Mechanics 2 Tutorial 9: The Lorentz Group Yehonatan Viernik December 27, 2020 1 Remarks on Lie Theory and Representations 1.1 Casimir elements and the Cartan subalgebra Let us first give loose definitions, and then discuss their implications. A Cartan subalgebra for semi-simple1 Lie algebras is an abelian subalgebra with the nice property that op- erators in the adjoint representation of the Cartan can be diagonalized simultaneously. Often it is claimed to be the largest commuting subalgebra, but this is actually slightly wrong. A correct version of this statement is that the Cartan is the maximal subalgebra that is both commuting and consisting of simultaneously diagonalizable operators in the adjoint. Next, a Casimir element is something that is constructed from the algebra, but is not actually part of the algebra. Its main property is that it commutes with all the generators of the algebra. The most commonly used one is the quadratic Casimir, which is typically just the sum of squares of the generators of the algebra 2 2 2 2 (e.g. L = Lx + Ly + Lz is a quadratic Casimir of so(3)). We now need to say what we mean by \square" 2 and \commutes", because the Lie algebra itself doesn't have these notions. For example, Lx is meaningless in terms of elements of so(3). As a matrix, once a representation is specified, it of course gains meaning, because matrices are endowed with multiplication. But the Lie algebra only has the Lie bracket and linear combinations to work with. So instead, the Casimir is living inside a \larger" algebra called the universal enveloping algebra, where the Lie bracket is realized as an explicit commutator.
    [Show full text]
  • Quivers and the Euclidean Group
    Contemporary Mathematics Volume 478, 2009 Quivers and the Euclidean group Alistair Savage Abstract. We show that the category of representations of the Euclidean group of orientation-preserving isometries of two-dimensional Euclidean space is equivalent to the category of representations of the preprojective algebra of type A∞. We also consider the moduli space of representations of the Euclidean group along with a set of generators. We show that these moduli spaces are quiver varieties of the type considered by Nakajima. Using these identifications, we prove various results about the representation theory of the Euclidean group. In particular, we prove it is of wild representation type but that if we impose certain restrictions on weight decompositions, we obtain only a finite number of indecomposable representations. 1. Introduction The Euclidean group E(n)=Rn SO(n) is the group of orientation-preserving isometries of n-dimensional Euclidean space. The study of these objects, at least for n =2, 3, predates even the concept of a group. In this paper we will focus on the Euclidean group E(2). Even in this case, much is still unknown about the representation theory. Since E(2) is solvable, all its finite-dimensional irreducible representations are one-dimensional. The finite-dimensional unitary representations, which are of inter- est in quantum mechanics, are completely reducible and thus isomorphic to direct sums of such one-dimensional representations. The infinite-dimensional unitary ir- reducible representations have received considerable attention (see [1, 3, 4]). There also exist finite-dimensional nonunitary indecomposable representations (which are not irreducible) and much less is known about these.
    [Show full text]
  • Chapter 1 GENERAL STRUCTURE and PROPERTIES
    Chapter 1 GENERAL STRUCTURE AND PROPERTIES 1.1 Introduction In this Chapter we would like to introduce the main de¯nitions and describe the main properties of groups, providing examples to illustrate them. The detailed discussion of representations is however demanded to later Chapters, and so is the treatment of Lie groups based on their relation with Lie algebras. We would also like to introduce several explicit groups, or classes of groups, which are often encountered in Physics (and not only). On the one hand, these \applications" should motivate the more abstract study of the general properties of groups; on the other hand, the knowledge of the more important and common explicit instances of groups is essential for developing an e®ective understanding of the subject beyond the purely formal level. 1.2 Some basic de¯nitions In this Section we give some essential de¯nitions, illustrating them with simple examples. 1.2.1 De¯nition of a group A group G is a set equipped with a binary operation , the group product, such that1 ¢ (i) the group product is associative, namely a; b; c G ; a (b c) = (a b) c ; (1.2.1) 8 2 ¢ ¢ ¢ ¢ (ii) there is in G an identity element e: e G such that a e = e a = a a G ; (1.2.2) 9 2 ¢ ¢ 8 2 (iii) each element a admits an inverse, which is usually denoted as a¡1: a G a¡1 G such that a a¡1 = a¡1 a = e : (1.2.3) 8 2 9 2 ¢ ¢ 1 Notice that the axioms (ii) and (iii) above are in fact redundant.
    [Show full text]
  • Symmetry Groups in Physics Part 2
    Symmetry groups in physics Part 2 George Jorjadze Free University of Tbilisi Zielona Gora - 25.01.2017 G.J.| Symmetry groups in physics Lecture 3 1/24 Contents • Lorentz group • Poincare group • Symplectic group • Lie group • Lie algebra • Lie algebra representation • sl(2; R) algebra • UIRs of su(2) algebra G.J.| Symmetry groups in physics Contents 2/24 Lorentz group Now we consider spacetime symmetry groups in relativistic mechanics. The spacetime geometry here is given by Minkowski space, which is a 4d space with coordinates xµ, µ = (0; 1; 2; 3), and the metric tensor gµν = diag(−1; 1; 1; 1). We set c = 1 (see Lecture 1) and x0 is associated with time. Lorentz group describes transformations of spacetime coordinates between inertial systems. Mathematically it is given by linear transformations µ µ µ ν x 7! x~ = Λ n x ; which preserves the metric structure of Minkowski space α β Λ µ gαβ Λ ν = gµν : The matrix form of these equations reads ΛT g L = g : This relation provides 10 independent equations, since g is symmetric. Hence, Lorentz group is 6 parametric. G.J.| Symmetry groups in physics Lorentz group 3/24 From the definition of Lorentz group follows that 0 2 i i det[Λ] = ±1 and (Λ 0) − Λ 0 Λ 0 = 1. Therefore, Lorentz group splits in four non-connected parts 0 0 1: (det[Λ] = 1; Λ 0 ≥ 1); 2: (det[Λ] = 1; Λ 0 ≤ 1), 0 0 3: (det[Λ] = −1; Λ 0 ≥ 1); 4: (det[Λ] = −1; Λ 0 ≤ 1). The matrices of the first part are called proper Lorentz transformations.
    [Show full text]