The su(n) Lie Algebra Haonan Liu December 2, 2022 Contents 1 The SU(n) Lie group 1.1 The SU(n) group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2 The SU(n) Lie group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 3 3 2 The su(n) Lie algebra 2.1 Lie algebra . . . . . . . . . . . . . . . . . 2.2 Basis, structure constants, and generators 2.2.1 Basis and generators of su(n) . . . 2.2.2 Structure constants of su(n) . . . . 2.3 States and operators . . . . . . . . . . . . 2.3.1 General definition . . . . . . . . . 2.3.2 Operators in Lie algebra . . . . . . 2.4 The Lie correspondence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 3 5 6 7 7 7 8 8 3 Representations of su(n) 3.1 Representation theory of Lie algebras . . . . . . . . 3.2 Examples of reps . . . . . . . . . . . . . . . . . . . 3.2.1 Trivial/zero rep . . . . . . . . . . . . . . . . 3.2.2 Defining/natural/standard/tautological rep 3.2.3 Adjoint/regular rep . . . . . . . . . . . . . 3.3 Morphism between reps . . . . . . . . . . . . . . . 3.3.1 Intertwining operator . . . . . . . . . . . . 3.3.2 Equivalent reps . . . . . . . . . . . . . . . . 3.3.3 Reps of su(n) and sl(n, C) . . . . . . . . . . 3.4 Operations on reps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 9 10 10 11 11 13 13 13 13 13 4 Structure theory of sl(n, C) and its reps 4.1 Simple and semisimple Lie algebras . . . 4.1.1 Ideals and commutativity . . . . 4.1.2 Simplicity and semisimplicity . . 4.1.3 Levi decomposition . . . . . . . . 4.2 Reducible and irreducible reps . . . . . . 4.2.1 Reducibility . . . . . . . . . . . . 4.2.2 Complete reducibility . . . . . . 4.3 Complete reducibility and symmetry . . 4.3.1 Schur’s lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 15 15 16 16 17 17 17 18 18 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4 4.5 4.6 4.3.2 Hamiltonian & first definition of symmetry . . . . 4.3.3 Unitary rep & the second definition of symmetry . 4.3.4 Multiplet . . . . . . . . . . . . . . . . . . . . . . . 4.3.5 Characters . . . . . . . . . . . . . . . . . . . . . . Reps of sl(n, C) . . . . . . . . . . . . . . . . . . . . . . . . 4.4.1 Reps of sl(2, C) . . . . . . . . . . . . . . . . . . . . 4.4.2 Cartan subalgebra . . . . . . . . . . . . . . . . . . 4.4.3 Root decomposition . . . . . . . . . . . . . . . . . 4.4.4 Cartan-Weyl basis . . . . . . . . . . . . . . . . . . 4.4.5 Root system . . . . . . . . . . . . . . . . . . . . . 4.4.6 Classification of semisimple Lie algebras . . . . . . 4.4.7 Weight decomposition . . . . . . . . . . . . . . . . 4.4.8 Reps of sl(n, C) . . . . . . . . . . . . . . . . . . . . 4.4.9 Casimir operators . . . . . . . . . . . . . . . . . . Direct product to direct sum . . . . . . . . . . . . . . . . 4.5.1 Symmetric and antisymmetric tensors . . . . . . . 4.5.2 Direct product into direct sum . . . . . . . . . . . 4.5.3 The Clebsch-Gordan coefficients (CG coefficients) Summary of properties of sl(n, C) and su(n) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 19 20 21 22 22 24 25 28 30 36 37 39 40 41 41 41 41 42 5 The su(2) Lie Algebra and its reps 5.1 Generators and structure constants . . . . . . . . 5.2 The Lie correspondence between SU(2) and su(2) 5.2.1 From SU(2) to su(2) . . . . . . . . . . . . 5.2.2 From su(2) to SU(2) . . . . . . . . . . . . 5.3 Reps . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.1 Defining rep . . . . . . . . . . . . . . . . . 5.3.2 Adjoint rep . . . . . . . . . . . . . . . . . 5.3.3 Irrep . . . . . . . . . . . . . . . . . . . . . 5.3.4 CG coefficients . . . . . . . . . . . . . . . 5.4 Cartan subalgebra . . . . . . . . . . . . . . . . . 5.4.1 Casimir operators . . . . . . . . . . . . . 5.4.2 Fundamental rep . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44 44 45 45 45 45 45 45 46 48 49 49 50 6 The su(3) Lie algebra and its reps 6.1 Generators and structure constants . . . . . . . 6.2 Reps . . . . . . . . . . . . . . . . . . . . . . . . 6.2.1 Fundamental and antifundamental reps 6.2.2 Tensor product rep . . . . . . . . . . . . 6.2.3 Multiplets . . . . . . . . . . . . . . . . . 6.2.4 Cartan subalgebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50 50 50 50 50 50 50 . . . . . . 7 The su(4) Lie algebra and its reps 51 8 Others 51 9 References 52 2 The SU(n) Lie group 1 The SU(n) group 1.1 1 . Definition 0 [U(n) group]: A unitary group of degree n, U(n), is the group of complex linear automorphisms of an n-dimensional complex vector space Cn preserving a positive definite Hermitian inner product H on Cn . In Def. 0, if the Hermitian form H is “standard”, meaning it is defined associated with the identity matrix 1 such that H(v, w) = v † · 1 · w, which is usually the case, then the group U(n) is just the group of n × n unitary matrices U , i.e., U † = U −1 . It can be proved that |det(U )| = 1 for any U ∈ U(n). Definition 0 [SU(n) group]: A special unitary group, SU(n), is the subgroup of U(n) with determinant 1. Easy to see, SU(m) is a subgroup of SU(n) for any m < n. Since SU(1) only contains the 1-dimensional identity, it is trivial and not interesting. Thus we will always have n > 1 for SU(n) in this work. The SU(n) Lie group 1.2 The SU(n) group is more than just a group, but a Lie group. Definition 0 [Lie group]: A Lie group is a set endowed simultaneously with the compatible structures of a group and a C ∞ (infinitely differentiable) manifold, such that both the multiplication and inverse operations in the group structure are smooth maps. A morphism between two Lie groups is just a map that is both differentiable and a group homomorphism. We have already defined SU(n) as a group. Multiplication is smooth because the matrix entries of the product U1 U2 are polynomials in the entries of U1 and U2 for U1 , U2 ∈ SU(n). Inversion is smooth by Cramer’s rule. Therefore SU(n) is a smooth manifold and thus a Lie group.2 One can go further and show that the SU(n) Lie group is compact, simply connected, and nonabelian.3 The benefit of understanding the Lie group nature of the SU(n) group is immediate once we introduce the one-to-one correspondence between Lie groups and Lie algebra in the next section. With the correspondence, most of the properties of the SU(n) group that we will use for our physical problems can be derived using the su(n) or sl(n, C) Lie algebra structure theory. Some useful properties of SU(n): 1. The matrix elements of an n × n matrix U ∈ SU(n) are analytic functions of d real parameters x1 , x2 , · · · , xd , with d = n2 − 1. This directly follows from the fact that SU(n) is smooth and that a unitary matrix with determinant 1 has n2 − 1 degrees of freedom in real numbers. Therefore we also call an SU(n) group a real Lie group of dimension n2 − 1. 2. More will be added as we complete this note. The su(n) Lie algebra 2 2.1 Lie algebra A Lie algebra4 is defined as the following. 1 All the ??? notations in the work need to be settled. All the "it can be shown" needs to be shown or cited. SU(n) ⊆ U(n) ⊆ GL(n, C) are all Lie groups and form smooth embeddings. 3 These properties are somewhat “nice” properties in representation theory. 4 Sometimes called a tangent space in literature, related to the smooth manifold nature of the corresponding Lie group. 2 Generally, 3 Definition 0 [Lie algebra]: A Lie algebra g is a vector space together with a skew-symmetric bilinear map 5 (called the Lie bracket) [ , ]:g×g→ − g (1) satisfying the Jacobi identity [a, [b, c]] + [b, [c, a]] + [c, [a, b]] = 0 (2) with a, b, c ∈ g. Note that, in this work, whenever we mention a Lie algebra, we will always assume that the Lie algebra is finite-dimensional over an algebraically closed 6 field F of characteristic 7 0 such as R or C unless explicitly noted. We can then define the Lie algebra su(n) as follows. Definition 0 [u(n), sl(n, C), and su(n) Lie algebras]: Let gl(n, C) be the Lie algebra that constitutes all the n × n complex matrices. The u(n) Lie algebra is a Lie algebra that constitutes all the anti-Hermitian n × n complex matrices, i.e., u(n) = U ∈ gl(n, C) | U † = −U . The sl(n, C) Lie algebra is a Lie algebra that constitutes all the traceless n × n complex matrices, i.e., sl(n, C) = {U ∈ gl(n, C) | Tr[U ] = 0}. The su(n) Lie algebra is a Lie algebra that constitutes all the traceless anti-Hermitian n × n matrices, i.e., su(n) = U ∈ gl(n, C) | Tr[U ] = 0 and U † = −U . For all these algebras, the skew-symmetric bilinear map is the commutator.8 Some comments on these algebras: 1. Notation. In this work, we write gl(n, C) = gl(Cn ). (3) More generally, we write gl(V ) to represent the set of automorphisms (matrices) from vector space V to V . 2. Both u(n) and su(n) are real Lie algebras. “Real” means that the algebra is a vector space over the field R, although the matrices themselves can contain complex numbers as entries. On the other hand, sl(n, C) is a complex Lie algebra. 3. Lie algebra su(n) is a subalgebra of u(n). Here, a subalgebra g0 of Lie algebra g is a subset of g of which the elements form a Lie algebra with the same commutator over the same field as those of g. 4. Complexification of a real Lie algebra. A real Lie algebra can be complexified if its basis elements stay linearly independent and the field R the Lie algebra is defined over is replaced by C. Both u(n) and su(n) satisfy this condition and can be complexified. Specifically, we define suC (n) ≡ su(n) ⊗ C = su(n) ⊕ isu(n), 5 “Skew-symmetric”, (4) or antisymmetric, means [X, Y ] = −[Y, X]. field F is algebraically closed if every polynomial with coefficients in F has a root in F 7 A field has characteristic 0 if the sum of its multiplicative identity never reaches its additive identity. 8 Definition (0) can also be phrased another way using the notion of exponential maps. The anti-Hermitian and traceless nature of matrices can be derived following rules of exponential maps and the definition of SU(n) Lie groups. However, following our logic in this work, we will define SU(n) and su(n) separately first, and introduce the Lie correspondence in the end together with the exponential map. 6A 4 Realizing that the Lie algebras su(n) and isu(n) are isomorphic and that they combined contain all the traceless n × n complex matrices, we have suC (n) = sl(n, C), (5) i.e., sl(n, C) is the complexification of su(n). It turns out that sl(n, C) will be very important in the theory. In fact, many of the properties of su(n) will be studied as a result of sl(n, C). 5. We mention Ado’s theorem here. Theorem 1 [Ado]: Any finite-dimensional Lie algebra over F is isomorphic to a Lie subalgebra of gl(n, F). Ado’s theorem tells us that without loss of generality Lie algebras can be studied with their elements defined as matrices. The su(n) Lie algebra is finite-dimensional. Another example of a finite-dimensional Lie algebra is the Heisenberg algebra which is spanned by the coordinate functions over R, i.e., {1, q, p}, with the skewsymmetric bilinear map defined as the Lie bracket. The Heisenberg algebra is useful in physics. See Miller or Woit for more. 2.2 Basis, structure constants, and generators We first discuss the properties of general Lie algebras. For a Lie algebra of dimension d over field F we can always find a basis of dimension d that we label as ej , j ∈ N. The Lie algebra is then denoted as spanF {e1 , e2 , · · · , ed } ≡ {e1 , e2 , · · · , ed }F . (6) Given a nonsingular matrix R, a new basis can be built using e0k = Rkj ej . Here, for simplicity, we have assumed the Einstein’s summation convention, and we do not care about the upper or lower indices of a tensor unless otherwise noted. From now on, we will also use the basis notation (6) to label a Lie algebra. If the field F is clear from context [R for su(n), C for sl(n, C), etc.] we will discard it as well. Applying the skew-symmetric bilinear map to basis elements yields another element of the Lie algebra that can be expanded in the basis, i.e., [ej , ek ] = fjkl el . (7) Here, the coefficients fjkl are called structure constants relative to the chosen basis. Several properties of the structure constants are: 1. Given a basis, the structure constants can describe the Lie algebra completely. 2. If the structure constants are all real, then the Lie algebra is real. 3. The structure constants transform between basis in the following way fj0 0 k0 l0 = Rj 0 j Rk0 k fjkl R−1 ll0 4. The structure constants are antisymmetric in the first two indices as a result of the Jacobi identity. We next focus on the su(n) algebra. 5 . 2.2.1 Basis and generators of su(n) By definition, the su(n) algebra is spanned by all the traceless anti-Hermitian matrices of dimension n × n. It follows by counting the real degrees of freedom that su(n), as a vector space, has a finite dimension d = n2 − 1. (8) Therefore, we can find a basis of dimension n2 − 1 for su(n), n > 2. The most common choice of basis in literature can be constructed in the following way. We first define a set {λi } as the set of the following traceless Hermitian matrices 1 0 1 0 0 1 0 0 0 0 1 s 1 0 1 0 −1 2 (9) {λi } = −2 . . ,··· , , , √ n(n − 1) 3 . 1 . 0 0 0 −(n − 1) 0 0 0 1 0 0 0 1 0 0 0 0 0 0 1 0 0 0 (10) . . . , 1 ,··· , , . . . 1 0 0 0 0 0 1 0 0 0 −i 0 0 0 0 0 −i 0 0 0 0 0 i 0 (11) . . . , i ,··· , . . . −i . 0 0 0 i 0 0 0 In more concise expressions, the set {λi } can be constructed from the following three classes of n×n matrices: s for µ < k 1 i h 2 (1) = λk (12) δµν × −k for µ = k, k = 1, 2, · · · , n − 1 ; k(k + 1) µν 0 for µ > k h i (2) λjk = δkµ δjν + δkν δjµ , j, k = 1, 2, · · · , n, and j < k; (13) µν h i (3) λjk = −i(δjµ δkν − δjν δkµ ), j, k = 1, 2, · · · , n, and j < k. (14) µν Here the superscripts (1), (2), and (3) represent the three classes. We can check that the total number of λi n(n − 1) n(n − 1) + = n2 − 1 = d. matrices is (n − 1) + 2 2 The matrices λi are called the generators of su(n).9 A basis of the su(n) is then just given by the set of anti-Hermitian traceless matrices {ei } defined from {λi } 1 ej = − iλj , 2 j = 1, 2, · · · , d. (15) The multiplication by i creates anti-Hermitian basis elements from the Hermitian generators. The coefficient 1 − is chosen for historical reasons. As mentioned above, our choice of generators and basis is not unique, 2 but has been shown convenient from literature. 9 Notice that the choice of generators is not unique, as is true for basis. Our choice is only one of the fundamental representations??? of su(n). 6 2.2.2 Structure constants of su(n) By the definition of structure constants in Eq. (7), using Eq. (15), we have [λj , λk ] = 2ifjkl λl . (16) With the choice of generators given by Eqs. (9)–(11) or Eqs. (12)–(14), we get a nice normalization condition for the generators λi , i.e., 10 Tr[λj λk ] = 2δjk (17) Now, using Eqs. (16) and (17) together, we get Tr [λj , λk ], λl = 2ifjkm Tr[λm λl ] = 2ifjkm 2δlm = 4ifjkl , (18) which gives the structure constants as fjkl = 1 Tr [λj , λk ]λl . 4i (19) On the other hand, using Tr[AB] = Tr[BA] in Eq. (18), we get 4ifjkl = Tr [λj , λk ]λl = Tr[λj λk λl − λk λj λl ] = − Tr[λj λl λk − λl λj λk ] = − Tr [λj , λl ]λk = −4ifjlk . (20) Similarly, we can prove that every odd permutation of the indices of fjkl changes its sign.11 In other words, fikl is completely antisymmetric in all indices. Therefore all structure constants 2 with the same indices vanish. n −1 To find all the nonvanishing terms of structure constants, we only need independent calculations, 3 given the completely antisymmetric property. For future use, we can also define the anticommutation relations for the generators λi . Without proof, for su(n), we have 4 + λl , (21) [λj , λk ]+ = δjk 1n×n + 2fjkl n where 1 + (22) fjkl = Tr [λj , λk ]+ λl . 4 2.3 States and operators 2.3.1 General definition We first discuss the concepts of states and linear operators in general. For an m-dimensional vector space over field F, we sometimes refer to the vectors as states or kets. The basis vectors 12 {ψ1 , ψ2 , · · · , ψm } can also be denoted as {|ψ1 i , |ψ2 i , · · · , |ψm i}. A linear operator is a morphism 13 that maps states from a vector space to another. For an operator Ω̂ : V → U , with {vj } and {uj } being bases of V and U , we can define the operator as k Ω̂vj = (Ωvj ) uk , (23) k where we have used the ˆ· notation to label operators. We call Ωkj = (Ωvj ) the matrix of the operator Ω̂. Ignoring the upper and lower indices (assuming Euclidean metrics for v and u), we get Ω̂vj = Ωkj uk , (24) where we have chosen the order of the lower indices on purpose. The reason will be explained shortly. 10 This actually why we chose this basis in the beginning. arbitrary Lie algebra, fjkl is only antisymmetric in the first two indices. For compact and semisimple Lie algebras such as su(n), fikl is completely antisymmetric, which is nice. 12 In this work we follow the notation in Eq. (6) such that a set of basis vectors can mean either the basis or the vector space spanned by the basis. The exact meanings should be clear from context. 13 A morphism is a structure-preserving map used in category theory. 11 For 7 2.3.2 Operators in Lie algebra Now we go back to the discussion on Lie algebra. So far we have seen that the Lie algebra elements of su(2) are matrices. Also we know that matrices are morphisms acting on certain vector spaces. Therefore, we can think of a Lie algebra g as a vector space consisting of linear operators that act on an m-dimensional vector space V . We assert that the operators that form the Lie algebra are morphisms from V to V . Let {v1 , v2 , · · · , vm } be a orthonormal basis of V , then we have, for Ω̂ ∈ g, using Eq. (24) Ω̂vj = Ωkj vk . (25) Furthermore, we demand the vector space V to be an inner product space. This allows us to define the inner product hvj |vk i, where hvj |’s are called bras that are a linear functional of kets living in the dual vector space V ∗ of V and satisfy † hvj | = |vj i . (26) In an inner product space, it is always possible to find orthonormal basis for the kets such that hvj |vk i = δjk , j, k ∈ {1, 2, · · · , m}. (27) Therefore we will always assume our basis of V and V ∗ to be orthonormal from now on. With orthonormal basis chosen, we can immediately appreciate our choice of the order of indices in Eq. (24). Realizing that different orders of indices are equivalent to matrix transpositions, our choice ensures the matrix Ωjk to have the nice property such that E D (28) hvk |Ω̂|vj i ≡ vk Ω̂vj = Ωlj hvk |vl i = Ωkj , where we have defined a left action. In other words, the elements of the matrix Ωkj are just the operator Ω̂ acted by bra hvk | on the left and ket |vj i on the right following the left action convention. Another nice property is that D E D ED E (ΩΩ0 )jk = hvj |Ω̂Ω̂0 |vk i = vj Ω̂Ω̂0 vk = vj Ω̂vl vl Ω̂0 vk = Ωjl Ω0lk , (29) i.e., the matrices of operators indeed follow our usual convention of matrix multiplications, which assumes left action. From now on, we will always think of elements of Lie algebra g as linear operators that act on some inner vector space V . 2.4 The Lie correspondence It is no coincidence that the SU(n) Lie group and su(n) Lie algebra have the same dimension d = n2 − 1. It can be proved that the linear SU(n) Lie group corresponds to the real su(n) Lie algebra, meaning that for every SU(n) Lie group, there exists a corresponding su(n) Lie algebra, and vice versa. This in general is called the Lie correspondence given by three fundamental theorems of Lie theory, and the one-to-one correspondence is only true for simply connected Lie groups, such as SU(n). Mathematically, the correspondence can be described using the following theorem. Theorem 2 [Lie correspondence]: The categories of finite-dimensional Lie algebras and connected, simply connected Lie groups are equivalent. 14 Since SU(n) is connected and simply connected, the categories of SU(n) and su(n) are equivalent, i.e., they correspond to each other. The Lie correspondence between SU(n) and su(n) have the following consequences: 14 I do not fully understand the equivalence of categories and will not go into details, but this is the correct statement. 8 1. Given a matrix U ∈ SU(n) parameterized by a set of real parameters {xj }, j ∈ {1, 2, · · · , d}, a basis of the corresponding su(n) Lie algebra can be found by (el )jk = ∂ Ujk (x1 , x2 , · · · , xd )|x1 =x2 =···=xd =0 . ∂xl (30) This relation is local in the parameter space. It directly relates the d = n2 − 1 degrees of freedom of SU(n) and su(n). From geometry point of view, su(n) is the tangent space to the Lie group SU(n) at the identity. 2. Given a n × n-dimensional basis {ej } of the su(n) Lie algebra, j ∈ {1, 2, · · · , d}, any matrix U ∈ SU(n) can be found with parameterization {xj }, j ∈ {1, 2, · · · , d}, by 15 i U (x) = exp(xj ej ) = exp − xj λj . (31) 2 This relation is nonlocal since it relies on the exponential map. Following Eq. (31), we can use the basis (or generators) of su(n) to reconstruct the group elements of SU(n). In other words, the generators of the su(n) Lie algebra are the infinitesimal generators of the SU(n) Lie group. In the next section we introduce the representation theory of Lie algebras. Due to the Lie correspondence, we will mainly focus on the su(n) Lie algebra. However, we will also introduce the representation theory of Lie groups in times of need. Representations of su(n) 3 3.1 Representation theory of Lie algebras Definition 0 [representation of Lie algebras]: A representation (or a rep) of a Lie algebra g over F on a vector space V (of dimension m) is a homomorphism ρ : g → gl(V ) = End(V ). Definition. 0 has the following consequences: 1. Notation. In many books the reps are defined to be the vector spaces they act on directly. In our notation, we will refer to ρ as the rep, and the vector space V as the g-module associated with the rep ρ of Lie algebra g. 2. Rep ρ is linear (by homomorphism), i.e., for α, β ∈ F and a, b ∈ g ρ(αa + βb) = αρ(a) + βρ(b). (32) 3. Rep ρ preserves the Lie brackets (by homomorphism), i.e., ρ([a, b]) = [ρ(a), ρ(b)]. (33) 4. Rep ρ can be written as m×m matrices if we specify the elements of the vector space V as m-dimensional vectors. Then we say rep ρ is m-dimensional. 5. Due to linearity, for a Lie algebra, it suffices to find the corresponding matrices of the basis to define a specific rep. 15 Surjective, proof not shown. 9 6. The operators we constructed in Sec. 2.3.2 are just another way to define reps. We will unify the concepts of operators and reps from now on. A Lie algebra g contains elements that satisfy the Jacobi identity. Once we say these elements are also linear operators acting on a vector space V , that automatically means that there is a homomorphism that maps these elements to the endomorphisms on V . As an add-on to our last point above, here n we show o that the set of operator matrices defined in Sec. 2.3.2 (Ωjk , etc.) is a rep of the Lie algebra g = Ω̂, · · · . Proof. The linearity is trivial to show. The commutation-preserving property can be shown as follows. Let Â, B̂ ∈ g be two operators of the Lie algebra. Then their operator matrices are just Ajk and Blm . Thus ([A, B])jk = (AB)jk − (BA)jk = Ajl Blk − Bjl Alk . What we have done above is more of a consistency check. We specify that for a Lie algebra composed of operators that act on the inner product space V , the corresponding operator matrices always form a rep of the Lie algebra. With the different inner product spaces chosen, the operator matrices have different matrix forms, i.e., the Lie algebra has different reps. From now on, for simplicity, we abandon the ˆ· operator notation. All the Lie algebra elements we talk about in this work will be taken as linear operators acting on the same vector space V , which is the g-module of the corresponding rep that maps these Lie algebra elements to the operators. Some other useful definitions regarding reps of Lie algebras. 1. Consider a rep ρ : g → gl(V ) of a Lie algebra g. A subspace U ⊂ V is invariant or stable if ∀u ∈ U and ∀a ∈ g, ρ(a)u ∈ U . Then we say ρ : g → gl(U ) is a subrepresentation (subrep). 2. A rep is faithful if it is injective. The rep of a Lie group is defined as below. Definition 0 [representation of Lie groups]: A rep of a Lie group G over F on a vector space V (of dimension m) is a homomorphism ρG : G → GL(V ) = Aut(V ). Here GL(V ) is the group of all m × m invertible matrices on the m-dimensional vector space V . In general, representations allow us to construct elements of the algebra or group as matrices acting on different spaces. In the cases of SU(n) and su(n), the one-to-one Lie correspondence between them is naturally inhibited by their reps. Therefore we will still focus on the reps of su(n) with comments on SU(n) reps when necessary. We next look at some examples of reps. 16 3.2 3.2.1 Examples of reps Trivial/zero rep The trivial/zero rep of a Lie algebra g is the rep that takes all elements of g to the zero linear map, i.e., ρ : g → gl(V ) (34) ρ(a) = 0. (35) such that for arbitrary V and all a ∈ g, Similarly, the trivial rep of a Lie group sends group elements to the identity matrix, so is also called the identity rep. 16 I was very overwhelmed by all these different concepts so I make sure to include all the names I can find. These concepts are not mutually exclusive. 10 3.2.2 Defining/natural/standard/tautological rep The defining/natural/standard/tautological rep is the rep for which the Lie algebra is naturally defined. For example, we have defined the su(n) Lie algebra as in Def. 0. Therefore, the defining rep is ρ : su(n) → gl(Cn ) (36) ρ(ej ) = (ej )kl (37) such that for j ∈ {1, 2, · · · , d} and k, l ∈ {1, 2, · · · , n}, where ej ’s are the basis elements of su(n) and (ej )kl ’s are their matrix forms on Cn given by Eqs. (12)–(14) and (15). The defining/natural/standard/tautological rep of a Lie group is defined likewise. For finite groups like S3 , the defining rep also means that there is a natural geometry interpretation. 3.2.3 Adjoint/regular rep The adjoint/regular rep of a Lie algebra g of dimension d is the rep for which the g-module is g itself, i.e., ad : g → gl(g) (38) ad(a)b = [a, b]. (39) such that for all a, b ∈ g, Clearly, the effect of the adjoint rep is to take the Lie bracket. We can define the adjoint operator of a ∈ g as ad(a) by ad(a)b = [a, b] = ad(a)jk bj , (40) where j, k ∈ {1, 2, · · · , d} since by definition the dimension of the adjoint matrices is d×d. Comparing Eqs. (7) and (40), we find that the adjoint matrices of basis elements are connected to the structure constants by ad(ej )lk el = fjkl el . (41) This is also the explicit form of the adjoint rep in basis {el }. The fact that the adjoint rep is a rep is equivalent to the Jacobi identity (2). Proof. Consider ad([a, b]) = ad(a)ad(b) − ad(b)ad(a) (42) ad([a, b])c = ad(a)ad(b)c − ad(b)ad(a)c [[a, b], c] = [a, [b, c]] − [b, [a, c]] [a, [b, c]] + [b, [c, a]] + [c, [a, b]] = 0. For a compact Lie group G and its corresponding Lie algebra g, the adjoint rep of the Lie group G is the homomorphism Ad : G → GL(g) such that for b ∈ g and A ∈ G Ad(A)b = AbA−1 ∈ g, (43) i.e., the adjoint operation is given by the conjugation of b by A. Let A = exa for x ∈ F and a ∈ g, then we have by BCH AbA−1 = exa be−xa = b + [a, b]x + O x2 . (44) 11 Therefore, using the Lie correspondence we have ad(a)b = [a, b] (45) which is exactly Eq. (39). The adjoint rep is useful in physics since it is closely related to invariant theory, which can be seen from its form in Eq. (43). 17 In Sec. 4.3.2 we will see how we can relate adjoint reps to symmetry through this reasoning. 3.2.3.1 Killing form Another useful tool constructed from the adjoint rep is the Killing form K(a, b). Given a rep ρ of g on V , a bilinear form 18 B on V is invariant under the action of g if B(ρ(a)v, w) + B(v, ρ(a)w) = 0 (46) for v, w ∈ V and a ∈ g. Now let ρ be the adjoint rep with V = g. Then we say a bilinear form B on g is invariant if B(ad(a)b, c) + B(b, ad(a)c) = 0 (47) B([a, b], c) + B(b, [a, c]) = 0 (48) or for a, b, c ∈ g. One important example of such invariant bilinear forms on g is the Killing form K(a, b), defined as K(a, b) = Tr [ad(a)ad(b)] (49) for a, b ∈ g. One can check that K(a, b) is also symmetric. For this reason, if the Killing form is also invertible, it can be interpreted as a metric tensor. This property will be used in Sec. 4.4. The Killing form of two basis elements of g, following Eq. (41), can be given by the structure constants, i.e., K(ej , ek ) = fjlm fkml . (50) Specifically for the su(n) Lie algebra, the Killing form can be further simplified and yields K(ej , ek ) = fjlm fkml = −fjlm fklm . (51) K(ej , ej ) = −fjlm fjlm , (52) Therefore the diagonal elements are which means Killing form on su(n) is negative definite. We say a Lie algebra is compact if its Killing form is negative definite. 19 A compact Lie algebra can be seen as the smallest real form of a complexification. In our case, su(n) is the real form of sl(n, C). We can also directly calculate the Killing form of su(n) using their defining reps. For x, y ∈ su(n), the Killing form is given by K(x, y) = 2n Tr[xy]. (53) Equation (53) also works for sl(n, C) although technically we have not given its defining rep (see Sec. 4.4.8). Killing forms are useful in the structure theory of Lie algebras in Sec. 4. 17 Is this correct??? bilinear form B(x, y) can have a coordinate representation as B(x, y) = xT Ay where A is the matrix of the bilinear form. This footnote is a self reminder and has nothing to do with the context. 19 There is some ambiguity here regarding whether the definition of a compact Lie algebra should correspond to a Killing form being negative definite or negative semidefinite. But su(n)’s Killing form is negative definite anyway so we do not care. 18 A 12 3.3 Morphism between reps In this section we discuss morphisms between reps. 3.3.1 Intertwining operator Definition 0 [intertwining operator]: An intertwining operator between two reps ρV and ρW of g (respectively, G) is a morphism f : V → W that commutes with the action of g (respectively, G), i.e., f ρV (a)v = ρW (a)f (v) (54) for a ∈ g (respectively, G) and v ∈ V . Notice that here we follow the convention that intertwining operators map one g-module to another g-module, instead of from rep ρV to rep ρW . The space of all g-morhisms between V and W is denoted as Hom(V, W ). 3.3.2 (55) Equivalent reps If f is bijective (isomorphism), then V and W are equivalent. The two reps ρV and ρW are then called equivalent reps. When two reps are equivalent reps, we say they are the same rep up to isomorphism, denoted by ∼ ρW . ρV = Equivalent reps can be transformed to one another using similarity transformations. 3.3.3 (56) 20 Reps of su(n) and sl(n, C) Theorem 3: Let g and gC be a real Lie algebra and its complexification. Then categories of complex reps of g and gC are equivalent. Theorem 3 means that any complex rep of g has a unique corresponding rep of gC , and Hom(V, W )g = Hom(V, W )gC . In other words, if we classify the reps of gC , the reps of g are also classified. In our case, Thm. 3 tells us that categories of complex reps of su(n) and sl(n, C) are equivalent. In fact, considering Thm. 2, the categories of finite-dimensional reps of SL(2, C), SU(2), sl(2, C), and su(2) are all equivalent. 21 3.4 Operations on reps Theorem 4: Let V and W be g-modules associated with reps ρV and ρW , respectively. Then there is a canonical structure 22 of a rep on V ⊕ W , V ⊗ W , and V ∗ . Specifically, the actions of g on these spaces are: 1. Action of g on V ⊕ W is given by ρV ⊕W : g → gl(V ⊕ W ) such that for a ∈ g, v ∈ V , and w ∈ W , ρV ⊕W (a)(v ⊕ w) = ρV (a)v ⊕ ρW (a)w. (57) 2. Action of g on V ⊗ W is given by ρV ⊗W : g → gl(V ⊗ W ) such that for a ∈ g, v ∈ V , and w ∈ W , ρV ⊗W (a)(v ⊗ w) = ρV (a)v ⊗ w + v ⊗ ρW (a)w. 20 A similarity transformation is diagonalization if the transformed matrix becomes diagonal. is true for n right??? 22 I think this means we can define a rep following the definition??? 21 This 13 (58) 3. Action of g on V ∗ is given by ρV ∗ : g → gl(V ∗ ) such that for a ∈ g, v ∈ V , and v ∗ ∈ V ∗ , ρV ∗ (a)v ∗ = −ρ∗V (a)v ∗ , (59) where ρ∗V (a) is the adjoint operator of ρV (a). Comments on Thm. 4: 1. The corresponding Lie group actions on V ⊕ W , V ⊗ W , and V ∗ for A ∈ G are given by ρV ⊕W (A)(v ⊕ w) = ρV (A)v ⊕ ρW (A)w, (60) ρV ⊗W (A)(v ⊗ w) = ρV (A)v ⊗ ρW (A)w, (61) ρV ∗ (A)v = ∗ ρ∗V (A −1 )v . ∗ (62) 2. The definition of ρV ⊗W is a natural result of linearity and the Lie correspondence. 3. The definition of ρV ∗ can be understood as follows. Consider the bilinear form h·, ·i which is given by h·, ·i : V × V ∗ → C, (63) where V ∗ is the dual space of V . Then following from the fact that the natural pairing V × V ∗ and the tensor product V ⊗ V ∗ have a universal property 23 , we can take V × V ∗ as the vector space V ⊗ V ∗ and understand the bilinear form h·, ·i as an intertwining operator between V ⊗ V and C (trivial rep). Using the definition of the intertwining operator and Eq. (58), we get hρV (a)v, v ∗ i + hv, ρV ∗ (a)v ∗ i = 0. (64) Equation (59) thus follows. Here the adjoint operator is the Hermitian adjoint or Hermitian conjugate operator in physics (usually labeled with “†”). In Dirac notation, Eq. (64) is just saying hv|ρV ∗ (a)ui = − hρV (a)v|ui E D hv|ρV ∗ (a)ui = − v ρ†V (a)u . (65) Thus we get Eq. (59). 4. Observing Eqs. (78) and (59), we see that ρV is a unitary rep if ρV = ρV ∗ . 5. As a direct result of Thm. 4, the following vector spaces are all g-modules if V and W are g-modules: (a) any tensor space V ⊗j ⊗ (V ∗ ) ⊗k for j, k ∈ N. (b) Hom(V, W ), with the action given by ρ(a)h = ρW (a)h − hρV (a) (66) for h ∈ Hom(V, W ). Specifically, when W = V , Hom(V, V ) = End(V ) ' V ⊗ V ∗ , i.e., we now have ρ : g → End(End(V )) = gl(gl(V )) = gl(V ⊗ V ∗ ) (67) ρ(a)h = ρV (a)h − hρV (a) (68) with where in Dirac notation h is just |vihu| for |vi ∈ V and hu| ∈ V ∗ . But Eq. (68) is just Eq. (39)! So we have recovered the adjoint rep. 23 See category theory. 14 Structure theory of sl(n, C) and its reps 4 We study the structure theory of Lie algebras with sl(n, C) as an example. With sl(n, C) being the complexification of su(n), all of the useful results regarding su(n) can be derived from those regarding sl(n, C). Mathematically, as we have mentioned, this means that categories of complex finite-dimensional reps of su(n) and sl(n, C) are equivalent. Therefore, with su(n) as our target, we investigate the classification theory of sl(n, C). We start from formalizing the simplicity of Lie algebra. 4.1 4.1.1 Simple and semisimple Lie algebras Ideals and commutativity The first property of the structure of a Lie algebra we discuss is its commutativity (or abelianity), i.e., how close a Lie algebra is to an abelian Lie algebra. The tool we use to study commutativity is called the ideal of a Lie algebra. A subalgebra g0 is invariant if [a, b] ∈ g0 , ∀a ∈ g0 and ∀b ∈ g. Then we define ideals as follows. Definition 0 [ideal]: An ideal i of a Lie algebra g is its invariant subalgebra, i.e., ∀g ∈ g, a ∈ i, [a, g] ∈ i. 24 Obviously, {0} and g are both ideals of g. Also, given i1 and i2 as ideals of g, i1 ∩ i2 , i1 + i2 , and [i1 , i2 ] are all ideals, where [i1 , i2 ] = {[a, b]} (69) for a ∈ i1 and b ∈ i2 . With the ideal defined, we now examine the commutativity of a Lie algebra. Specifically, we make the following definitions given a Lie algebra g: 1. The commutant of g is the ideal [g, g]. The smaller the commutant is, the closer g is to being abelian. We say that a Lie algebra g is abelian if [a, b] = 0, ∀a, b ∈ g. Then g is abelian when [g, g] = 0. 2. Lie algebra g is solvable if Dn g = 0 for some n ∈ N, where ( j+1 D g = Dj g, Dj g D0 g = [g, g] . (70) 3. A radical rad(g) is the unique solvable ideal that contains any other solvable ideal of g. 4. Lie algebra g is nilpotent if Dn g = 0 for some n ∈ N, where ( Dj+1 g = [g, Dj g] D0 g = g . (71) Both nilpotent and solvable Lie algebras are “almost abelian” algebras, meaning that an abelian algebra can be made by successive extensions of abelian algebras. As an example, let b ⊂ gl(n, F) be the subalgebra of upper triangular matrices and n ⊂ gl(n, F) be the subalgebra of all strictly upper triangular matrices. Then b is solvable, and n is nilpotent. We mention here there is a Cartan criterion for solvability, i.e., g is solvable iff K([g, g], g) = 0. 24 Since the commutator is skew-symmetric, the left ideal and right ideal are the same. 15 4.1.2 Simplicity and semisimplicity We next introduce semisimplicity and simplicity to describe how far a Lie algebra is from being abelian. Definition 0 [semisimplicity of Lie algebras]: A Lie algebra g is semisimple if any of the following equivalent statements is true: 1. Lie algebra g does not have a nonzero abelian ideal. 2. Lie algebra g does not have a nonzero solvable ideal. 3. The radical rad(g) = 0. 4. (Cartan criterion) The Killing form K(a, b) for a, b ∈ g is nondegenerate, i.e., det [K(a, b)] 6= 0. 25 Definition 0 [simplicity of Lie algebras]: A Lie algebra g is simple if it is not abelian and does not possess an ideal other than {0} and g. Here we give several useful criteria for semisimplicity of a Lie algebra: 1. A Lie algebra is semisimple iff it is a direct sum of simple Lie algebras. 2. Any compact Lie algebra is semisimple. 3. If a Lie algebra is simple, then it is semisimple. 4. gC is semisimple iff g is semisimple, where g and gC are a Lie algebra and its complexification. We will see in Sec. 4.2 that semisimple Lie algebras are completely reducible, which is a very desired property in physics. Semisimplicity means far away from commutativity. If a Lie algebra g is semisimple, then its commutant [g, g] = g. All of the classical Lie algebras (also the most useful Lie algebras in physics) [see Sec. 4.4.6] except for D1 and D2 are simple. 26 Since sl(n, C) is a classical Lie algebra, it is simple, thus semisimple. Therefore su(n) is semisimple. One counterexample is the Heisenberg algebra {X, P, 1} that is neither simple nor semisimple. 4.1.3 Levi decomposition Given the tools to describe a Lie algebra’s commutativity, we have the Levi theorem. Theorem 5 [Levi decomposition]: Any Lie algebra g can be written as a direct sum g = rad(g) ⊕ gss (72) where gss is a semisimple Lie algebra (not necessarily an ideal). The Levi theorem states that any Lie algebra can be written as a direct sum of an abelian part and a semisimple part. We can add more structure to the abelian part. If rad(g) = z(g), where z(g) = {a ∈ g | [a, b] = 0, ∀b ∈ g} is the center of g, then we say g is reductive. 27 Any simple Lie algebra is reductive. 25 Recall that a symmetric bilinear form on a finite-dimensional vector space V is nondegenerate/invertible if for all a 6= 0 there exists b such that K(a, b) 6= 0. 26 What is D ??? 1 27 Reductivity is a nice property for Lie algebras since it ensures [g, rad(g)] = 0, i.e., elements in an irrep will act as zero only when they are zeros. 16 4.2 4.2.1 Reducible and irreducible reps Reducibility We now turn from Lie algebras to their reps. We make the following definitions. Definition 0 [reducibility of reps]: A nontrivial rep ρ : g → gl(V ) of a Lie algebra g is irreducible if it has no subreps other than the trivial rep or itself. A rep is reducible if it is not irreducible. For convenience, we say the V is an irreducible sl(2, C)-module, or V is irreducible, if rep ρ : g → gl(V ) is irreducible. As an example, we have the following theorem. Theorem 6: The rep ρ : sl(n, C) → gl(Cn ) is irreducible. Theorem. 6 also means the defining reps of sl(n, C) are irreducible. Following Thm. (6), we clarify the difference between the direct sum of vector spaces Cl and Cm and the direct of sum of g-modules Cl and Cm . For arbitrary vector spaces Cl and Cm , we have Cl+m = Cl ⊕ Cm . However, for vector spaces Cl and Cm acted upon by reps of g, we have Cl+m 6= Cl ⊕ Cm (g-modules). (73) u with a ∈ g, u ∈ Cl , and This is because for rep ρCl+m : g → gl Cl+m , we require action ρCl+m (a) v v ∈ Cm . However for reps ρCl : g → gl Cl and ρCm : g → gl(Cm ), following Eq. (57), the rep of their direct u 0 u + ρCm (a) . Clearly, by definition ρCl ⊕Cm is reducible, while ρCl+m = ρCl (a) sum acts as ρCl ⊕Cm (a) 0 v v can be irreducible. In this sense, we have Eq. (73). Therefore Thm. 6 tells us that any rep of sl(n, C) on Cn is irreducible, n ∈ Z and n ≥ 2. Nevertheless any rep of sl(n, C) on Cm ⊕ Cn−m is obviously reducible, for m, n − m ∈ Z+ . 4.2.2 Complete reducibility Irreducible reps, or irreps, of semisimple Lie algebras have nice properties that they can be classified completely as we will see later. In fact, one of the most commonly used technique in physics is to decompose a direct product of irreps as a direct sum of irreps, which will then be useful due to their block diagonal nature. We formalize this process by introducing the complete reducibility of reps. Definition 0 [complete reducibility of reps]: A rep of g acting on V is completely reducible if it is isomorphic to a direct sum of irreps, i.e., V ' ⊕nj Vj for nj ∈ N. Here nj ’s are called multiplicities, and Vj ’s are gmodules of pairwise nonisomorphic reps. For convenience, we say V is a completely reducible g-module, or V is completely reducible, if the rep on V is completely reducible. Not every rep of a Lie algebra is completely reducible. Mathematically, we want to answer the following three questions: 1. For what Lie algebras are all the reps completely reducible? 2. For a given completely reducible rep of a Lie algebra, what is the decomposition into irreps? 3. For a given Lie algebra, classify all irreps. In this section, we answer the first question. The second question will be answered theoretically use characters in Sec. 4.3.5, but a practical method using weight decomposition will be given in Sec. 4.4. The third question will be briefly answered in Sec. 4.4.6. The answer to the first question is the following theorem. 17 Theorem 7: Any complex finite-dimensional rep of a semisimple Lie algebra is completely reducible. Since sl(n, C) is semisimple, any complex finite-dimensional rep of sl(n, C) is completely reducible, and so is that of su(n). In Sec. (4.4), we will focus on the decomposition of reps of semisimple Lie algebras, with sl(n, C) as an example. In the next section, we examine other theorems that are related to complete reducibility. This is a deviation from purely Lie algebra structure theory since we discuss compact Lie groups and their associated real Lie algebras such as SU(n) and su(n). However, the contents are very important in physics. 4.3 4.3.1 Complete reducibility and symmetry Schur’s lemma One useful tool to decompose completely reducible reps into irreps is the use of intertwining operators. Specifically, there is Schur’s lemma. Theorem 8 [Schur’s lemma]: Let ρV and ρW be complex irreps of g (respectively, G). 1. If ρV ∼ = ρW (V = W ), then Hom(V, W ) = Hom(V, V ) = C1. 2. If ρV ρW , then Hom(V, W ) = 0. Schur’s lemma tells us how inequivalent irreps are fundamentally different from each other. Part 1 says equivalent irreps only differ by coefficients. Part 2 says inequivalent irreps cannot be intertwined unless trivially. The following theorems are direct results of Schur’s lemma. Theorem 9: If g (respectively, G) is commutative, then any complex irrep of g (respectively, G) is onedimensional. Theorem 10: Let ρV be a completely reducible rep of g (respectively, G). L 1. If V = Vj , ρVj irreducible and pairwise nonisomorphic, then any intertwining operator Φ : V → V L is of the form Φ = λj 1Vj . L L nj 2. If V = nj Vj = C ⊗ Vj , ρVj irreducible and pairwise nonisomorphic, then any intertwining L operator Φ : V → V is of the form Φ = Aj ⊗ 1Vj , for Aj ∈ End(Cnj ). Theorem 10 gives us a very effective way to analyze intertwining operators. Specifically, it is extremely useful in physics as the Hamiltonian can be thought of as an intertwining operator. 4.3.2 Hamiltonian & first definition of symmetry Consider a Hilbert space V , which is a complex vector space, and a Hamiltonian operator H : V → V . Then the following statements are equivalent: 1. There is a symmetry described by a Lie group G. 2. There is an action of G on V that leaves H invariant. 3. For all A ∈ G, AHA−1 = H. 4. H is an intertwining operator that commutes with actions of G. 18 (74) The above statements can be taken as a first definition of symmetry. Specifically, it predicts a very nice physics picture as illustrated below. Given H as an intertwining operator H : V → V , if the Lie group G has a completely reducible rep on L V such that V = nj Vj is the G-module, then immediately from Thm. 10 we see that H has the form ··· c11 1Vj · · · c1nj 1Vj M (75) H= Aj ⊗ 1Vj = ··· ··· ··· . cnj 1 1Vj · · · cnj nj 1Vj ··· If all the nj ’s are 1, then H= M Ej 1Vj ··· = (76) . Ej 1Vj ··· Since H has a block diagonal form, the following statements are true: 1. If the H commutes with a rep of G, then the energy eigenspaces are irreps. Group actions of G on an energy eigenstate will not change its energy. 2. A state that lives in the vector space of an irrep will stay there under time evolution. 28 All this seems nice, only if G has a completely reducible rep. 4.3.3 Unitary rep & the second definition of symmetry It turns out that in our case G does have a completely reducible rep. To see this we first introduce the notion of unitary reps. Definition 0 [unitary rep]: A complex rep ρ : g → gl(V ) of a real Lie algebra g is unitary if there is an inner product that is g-invariant, i.e., for all a ∈ g and v, w ∈ V (ρ(a)v, w) + (v, ρ(a)w) = 0, (77) ρ(a)† = −ρ(a). (78) or equivalently, if ρ(a) ∈ u(V ), i.e., A complex rep ρG : G → GL(V ) of a real Lie group G is unitary if there is an inner product that is G-invariant, i.e., for all A ∈ G and v, w ∈ V , (ρG (A)v, ρG (A)w) = (v, w), (79) ρG (A)† = ρG (A)−1 . (80) or equivalently, if ρG (A) ∈ U(V ), i.e., Unitary reps are nice because they are completely reducible. Theorem 11: Every unitary rep is completely reducible. Therefore if we can show that Lie groups have unitary reps then we are done. This is given by the following theorem. 28 We will discuss this vector space, the multiplet, in the next section. 19 Theorem 12 [Peter-Weyl, Part II]: Any finite-dimensional rep of a compact Lie group G (respectively, g) is unitary and thus completely reducible. Notice that Thm. 12 discusses a narrower class of Lie algebras than Thm. 7, since compact Lie algebras [e.g., su(n)] are the compact real forms of the corresponding complex semisimple Lie algebras [e.g., sl(n, C)]. However, Thm. 12 does provide the unitarity property that Thm. 7 does not guarantee. From Thm. 12 we get the following statement. Theorem 13: Any finite-dimensional rep of SU(n) and su(n) is unitary and completely reducible. We can now collect our thoughts. Given a compact Lie group G with its associated real Lie algebra g and the complexification gC , we know that any finite-dimensional rep of G, g, and gC is completely reducible. Therefore, given a Hamiltonian operator, which is also an intertwining operator H : V → V , all the results from Sec. 4.3.2 will stand. This is nice, but where did the unitarity of reps of G and g play a role? Recall in Eq. (43) we have introduced the adjoint rep of a Lie group that leaves a Lie algebra element invariant, i.e., for a ∈ g and A ∈ G, AaA−1 ∈ g. (81) Now that we know any finite-dimensional rep of G or g is unitary from Thm. 12, we can define the unitary reps ρG : G → U(V ) and ρ : g → u(V ). Then for A ∈ G and a ∈ g we have ρG (A)ρ(a)ρG (A)−1 ∈ u(V ). (82) Let us look at Eq. (82) and think about symmetry again. If states in a Hilbert space transform under a Lie group action that is unitary, then we say the group is a group of symmetry of the Hilbert space since the group action preserves the inner product of states. Then observables, as linear operators on the Hilbert space, are well-defined if their transformation under the group action follows the transformation of the states in such a way that if the observables and states are transformed together then no information should be able to get measured. Since states transform with ρG (A), the observables should transform by a conjugation of these unitary reps, i.e., exactly given by Eq. (82). Considering Hermitianity, a Hermitian operator (observable) acting on V can thus be constructed as iρ(a), which are the generators of the Lie algebra. From Eq. (82), after transformation under the group action, the observables remain Hermitian. In this sense, we give the second definition of symmetry in quantum mechanics. Definition 0 [symmetry]: We say that there is a symmetry when there is a unitary rep of a Lie group acting on the Hilbert space. The Lie algebra generators both generate the unitary Lie group rep, and also transform under it as observables via conjugation. Normally, observables transform into other generators of the Lie algebra. Specifically, when the observable is the Hamiltonian H, and the group action leaves H invariant, we retain the first definition of symmetry in Sec. 4.3.2. 4.3.4 Multiplet L In Sec. 4.3.2 we see that for a completely reducible rep, its g-module can be written as V = nj V j . A state that lives in one of the g-modules of the irreps Vj remains in Vj when acted by the intertwining operator H. These g-modules of the irreps are called multiplets in physics. Definition 0 [multiplet]: The g-module of an irrep ρ of Lie algebra g is called the multiplet of ρ. The concept of multiplets appears everywhere in quantum physics. Here are some important properties: 1. Stability. Let ρ be a rep. If ρ is irreducible, then by definition states transform in the same multiplet under Lie algebra elements. If ρ is completely reducible, then due to the block diagonal form, states 20 that are in a certain multiplet will transform within the same multiplet under actions of Lie algebra elements. 2. State multiplets. We refer to the multiplet expressed in a specific state basis {vj } as the state multiplet associated with the basis. 3. The su(n) and sl(n, C) Lie algebras have the same multiplets, as a result of Thm. 3. Next we define the fundamental multiplets of the su(n) algebras. 29 Definition 0 [fundamental multiplet]: If the sl(n, C)-module is Cn , then we call this sl(n, C)-module the fundamental multiplet of sl(n, C). Similarly for su(n). By Thm. 6, the fundamental multiplet of of sl(n, C) is well-defined. The fundamental multiplet of sl(n, C) is the defining multiplet of sl(n, C), as well as the multiplet of lowest dimension corresponding to the irreps of sl(n, C). 4.3.5 Characters So far we have seen that a large class of reps are completely reducible. However, we have not talked about the decomposition of completely reducible reps or the multiplicity yet. In this section we introduce the concept of characters. Definition 0 [characters of real compact Lie groups and algebras]: Let G and g be a real compact Lie group and its associated Lie algebra. The character XV of a finite-dimensional rep ρG : G → GL(V ) is the function on G defined by XV (A) = Tr[ρG (A)]. The character χV of a finite-dimensional rep ρ : g → gl(V ) is the function on g defined by h i χV (a) = Tr eρ(a) . (83) (84) Definition .0 has the following consequences: 1. Characters are basis free. 2. The characters are connected by χV (a) = XV (ea ). (85) Since the characters of G and g are defined to be equal for corresponding elements, below we only discuss XV . 3. If ρG is trivial rep, then XV (A) ≡ 1. 4. For A, B ∈ G, XV ⊕W = XV + XW , (86) XV ⊗W = XV XW , (87) XV ∗ = X(ABA 29 I −1 XV∗ , (88) ) = XV (B). (89) do not know if this definition is useful at all. See Greiner for more on multiplets. 21 Characters offer a theoretical way to determine the multiplicities through an inner product on C ∞ (G, C) given by ˆ (f1 , f2 ) = f1 (A)f2∗ (A) dA (90) G where dA is the Haar measure. From Schur’s lemma in Thm. (8), we can prove Theorem 14: Let ρV and ρW be nonisomorphic complex irreps of a compact Lie group G. Then 1. (XV , XW ) = 0, which means the characters XV and XW are orthogonal with respect to the inner product. 2. (XV , XV ) = 1. This implies the following theorem. Theorem 15: Let ρV be a complex rep of a compact real Lie group G. Then 1. ρV is irrep iff (XV , XV ) = 1. L 2. V can be uniquely written in the form V = nj Vj , where Vj ’s are G-modules of pairwise nonisomor phic irreps and the multiplicities nj are given by nj = XV , XVj . Theorem. 15 gives a way to calculate nj . In reality, it is usable only for finite groups and special cases of Lie groups due to the difficulty to calculate the inner product (90). In Sec. 4.4 we will develop a much more practical method using weight decomposition. 4.4 Reps of sl(n, C) In this section we answer the second question we asked in Sec. 4.2, i.e., how to find the decomposition of a completely reducible rep. We will assume g to be a finite-dimensional complex semisimple Lie algebra for the remainder of Sec. 4. Specifically, we use sl(n, C) as an example, which is also our primary goal. For simplicity, whenever we specify the g-module, we will write ρ(a)v = av for a ∈ g, i.e., we treat a not only as the Lie algebra element, but also as an operator. This notation is closer to the one used in physics. We start from the simplest case of sl(n, C), the sl(2, C) Lie algebra. 4.4.1 Reps of sl(2, C) One most commonly used basis of sl(2, C) is {e, f, h} that satisfies [h, e] = 2e, [h, f ] = −2f, [e, f ] = h. In the defining rep, which is on C2 , these basis elements are commonly chosen as 0 1 0 0 1 0 e = σ+ = , f = σ− = , h = σz = . 0 0 1 0 0 −1 (91) (92) Notice that h is diagonal and has eigenvalues ±1. More generally, in an arbitrary rep, we know by the fundamental theorem of algebra that h can be diagonalized with complex eigenvalues. This can be formalized as the weight decomposition. 22 Definition 0 [weight, sl(2, C)]: Let V be an sl(2, C)-module. A vector v ∈ V is called a vector of weight λ, λ ∈ C, if it is an eigenvector of h with eigenvalue λ, i.e., hv = λv ⊂ Vλ , (93) where the subspace with eigenvalue λ is denoted as Vλ ⊂ V . Then the actions of e and f on Vλ can be derived from commutation relations (91), yielding ev ⊂ Vλ+2 , (94) f v ⊂ Vλ−2 . (95) Since e brings a vector from Vλ to Vλ+2 , operator e is usually called the raising operator. Similarly, since f brings a vector from Vλ to Vλ−2 , operator f is usually called the lowering operator. We can thus apply the weight decomposition to reps of sl(2, C). Theorem 16 [weight decomposition, sl(2, C)]: Every finite-dimensional rep of sl(2, C) has a weight decomposition, i.e., M V = Vλ . (96) λ By Thm. 7, any complex finite-dimensional rep of sl(2, C) is completely reducible. Therefore, it suffices to discuss only the weight decomposition of irreps. Specifically, we define the highest weight of V . Definition 0 [highest weight]: Let V be irreducible. A weight λ of V (Vλ 6= 0) is the highest weight if for every weight λ0 of V we have Re λ ≥ Re λ0 . (97) We can now write down the classification of irreps of sl(2, C). Theorem 17 [classification, irreps of sl(2, C)]: Let {h, e, f } be the basis of sl(2, C). 1. For any n ∈ N, let Vn be a (n + 1)-dimensional vector space with basis {v0 , v1 , · · · , vn }. Define the action of sl(2, C) by hvk = (n − 2k)vk , ( f vk = (k + 1)vk+1 , (98) k < n; f vn = 0, ( evk = (n + 1 − k)vk−1 , k > 0; ev0 = 0, (99) (100) Then ρVn : sl(2, C) → gl(Vn ) is an irrep. We call this irrep the irrep with highest weight n. 2. For n 6= m, reps ρVn and ρVm are nonisomorphic. 3. Every finite-dimensional irrep of sl(2, C) is isomorphic to one of the reps ρVn . See Fig. 1 for a figure presentation of Thm. 17. The classification of irreps of sl(2, C) provides the fundamental intuition for the classification of irreps of semisimple Lie algebras. To generalize this weight decomposition to sl(n, C) and other semisimple Lie algebras, we require the concept of the Cartan subalgebra. 23 1 vn n−2k −n+2 −n n 2 v n−1 n+1−k n−k vk ··· n−1 n−2 k+1 n n n−1 v1 ··· 2 k v0 1 Figure 1: Action of sl(2, C) on Vn in the irrep ρVn . Operator h sends a vector vk back to itself via the self loop arrows. Operators e and f act as the top and bottom arrows, respectively. 4.4.2 Cartan subalgebra For a finite-dimensional complex semisimple Lie algebra, the Cartan subalgebra can be defined as follows. Definition 0 [Cartan subalgebra of complex semisimple Lie algebras]: A Cartan subalgebra h is the maximal abelian subalgebra of Lie algebra g consisting semisimple elements. Semisimple elements are elements x ∈ g such that the adjoint operator ad(x) : g → g is semisimple (diagonalizable). In short, the Cartan subalgebra is the maximal abelian subalgebra with diagonalizable elements in the adjoint rep. In every complex semisimple Lie algebra, there exists a unique Cartan subalgebra up to isomorphism. The dimension l of this unique Cartan subalgebra h is called the rank of a Lie algebra. Definition 0 [rank]: The rank l of g is the dimension of the Cartan subalgebra h, i.e., l = dim h. (101) For sl(n, C), h= diag{c1 , c2 , · · · , cn } | cj ∈ C, X j cj = 0 (102) is a Cartan subalgebra. This will be the most important Cartan subalgebra we use in this work. The rank of g is the largest number of commuting basis elements (generators) if they are all diagonalizable in the adjoint rep. For example, the rank of sl(n, C) is l = n − 1, following from Eq. (102). Similarly, the rank of su(n) is also l = n − 1. The above definitions are enough for our use of Cartan subalgebras in physics. The more general definition of a Cartan subalgebra is rather subtle. Below we add a brief compendium of the full construction of the Cartan subalgebra. We first introduce the following concepts: 1. Normalizer. The normalizer of a subset s in a Lie algebra g is Ng (s) = {x ∈ g | [x, s] ∈ s, ∀s ∈ s}. (103) A subalgebra h of a Lie algebra g is self-normalizing if Ng (h) = h. 2. Centralizer. The centralizer of a subset s in a Lie algebra g is Cg (s) = {x ∈ g | [x, s] = 0, ∀s ∈ s}. (104) An element a ∈ g is regular if the dimension of its centralizer (if we take the element as the subset) is minimal among all centralizers of elements of g. Consider a finite-dimensional Lie algebra g over a field F. Then we have the following definition. Definition 0 [Cartan subalgebra, general]: A Cartan subalgebra is a nilpotent subalgebra h of a Lie algebra g that is self-normalizing. 30 30 In Kirillov there is another way to define Cartan subalgebra from its toral property. In Greiner and Muller P440 there is yet another way using simple linear algebras. 24 Specifically, 1. if F is infinite, then g has Cartan subalgebras; 2. if F is of characteristic zero, then there is a Cartan subalgebra for every regular element of g such that each regular element belongs to one and only one Cartan subalgebra and all Cartan subalgebras have the same dimension l; 3. if F is algebraically closed, then all Cartan subalgebras are isomorphic; 4. if g is semisimple, then all Cartan subalgebras are Abelian. Therefore, for semisimple Lie algebras over R or C, Def. 0 follows from Def. 0 and the properties above. For sl(2, C), the Cartan subalgebra is just Ch, which is one-dimensional. Theorem. 17 gives a full classification of all the irreps of sl(2, C) using the Cartan subalgebra and its corresponding raising and lowering operators. For a more complicated semisimple Lie algebra g, its turns out that we can always consider g as a module over sl(2, C) and use Thm. 17 to classify its reps. However, to understand this result we need to work with a special decomposition of g, the root decomposition, which is the weight decomposition in the adjoint rep. 4.4.3 Root decomposition Consider a semisimple Lie algebra g [such as sl(n, C)] and its Cartan subalgebra h as defined in Def. 0 which is unique up to isomorphism. Let {Hj }, j ∈ {1, 2, · · · , l}, be a basis of h ⊂ g. By definition, all Hj ’s commute with each other and can be simultaneously diagonalized in the adjoint rep. This means that the simultaneous eigenvectors of Hj ’s must span g, thus forming a basis. The eigenvectors with zero eigenvalues span h. Therefore, without loss of generality, we can choose these eigenvectors as Hj themselves. The eigenvectors with nonzero eigenvalues span g/h. For each of these eigenvectors, the set of l nonzero eigenvalues associated with the Hj ’s constitutes another vector of dimension l, called a root. Denoting such an eigenvector as Eα and its associated root as α, we can express the eigendecomposition of Hj by [Hj , Eα ] = αj Eα . (105) The form of Eq. (105) implies duality. We can take the root α as a linear functional on h by considering some bilinear form h·, ·i. In other words, all the roots together with zero span h∗ , the dual space of h. For convenience we define E0 ∈ h. However, when we refer to roots we always assume they are nonzero. We thus have the root decomposition theorem. Theorem 18 [root decomposition]: The root decomposition of g is ! M g=h⊕ gα (106) α∈R where α’s are called the roots, the subspaces gα = {x ∈ g | [h, x] = hα, hi x, ∀h ∈ h} (107) are called the root subspaces, and the set R = {α ∈ h∗ \{0} | gα 6= 0} (108) g0 = h. (109) is called the root system. Specifically, we define 25 The root decomposition is the weight decomposition in the adjoint rep. As an example, consider g = sl(n, C) with h defined in Eq. (102). Define ej : h → C as the functional h1 .. ej : (110) → hj . . hn P Since j ej = 0, we have , M X h∗ = Cej C ej . (111) j j Notice that matrix units Ejk are eigenvectors of ad(h) for h ∈ h since ad(h)Ejk = [h, Ejk ] = (hj − hk )Ejk = hej − ek , hi Ejk . (112) Thus the root decomposition of sl(n, C) is given by R = {ej − ek | j 6= k} ⊂ h∗ , gej −ek = CEjk . (113) (114) With the root decomposition, we can show the following properties: 1. For root subspaces, [gα , gβ ] ⊂ gα+β . (115) 2. If α + β 6= 0 then gα and gβ are orthogonal with respect to the Killing form K. 3. For any α, the Killing form K gives a non-degenerate pairing gα ⊗ g−α → C. (116) In particular, restriction of K to h is non-degenerate. So far we have not specified the natural pairing h·, ·i introduced in Eq. (107). One standard procedure to define this natural pairing is to consider a nondegenerate symmetric bilinear form (·, ·), i.e., a metric, defined on g. Such a metric exists. For example, by the Cartan criterion in Def. 0 and the last property above, the Killing form is one such metric for semisimple Lie algebras. In fact, if we take the metric as the Killing form, as we will see later, we can define the so-called Cartan-Weyl basis for a semisimple Lie algebra. However, for now, we just assume there exists such a metric (·, ·). Since the restriction of (·, ·) to the Cartan subalgebra h is nondegenerate, we can also define a metric on its dual h∗ , which we also denote as (·, ·). Finally, denoting the dual element of root α ∈ h∗ as Hα ∈ h, we make the following definition hα, Hβ i = (Hα , Hβ ) = (α, β), (117) for α, β ∈ h∗ . Immediately we can show that 1. Let e ∈ gα , f ∈ g−α . Then [e, f ] = (e, f )Hα . (118) (α, α) = (Hα , Hα ) 6= 0. (119) 2. For α ∈ R, 26 3. Let e ∈ gα , f ∈ g−α be such that 2 , (α, α) (e, f ) = (120) and let hα = 2Hα . (α, α) (121) Then hhα , αi = 2 and the elements e, f, hα satisfy the commutation relations (91) of sl(2, C). We denote such a subalgebra by sl(2, C)α ⊂ g. Moreover, the so defined hα is independent of the choice of the metric (·, ·). 4. The vector space M V = Chα ⊕ gkα ⊂ g (122) k∈Z,k6=0 is an sl(2, C)α -module. From the last statement (lemma) we see that the root decomposition offers a way to study the structure of any semisimple Lie algebra g by decomposing it into sl(2, C)α ’s and then use the theory of reps of sl(2, C) given in Sec. 4.4.1 to study it. The main theorem is given below. Theorem 19 [structure of semisimple Lie algebras]: Let g be a complex semisimple Lie algebra with Cartan subalgebra h and root decomposition given in Eq. (106). Let (·, ·) be a nondegenerate symmetric invariant bilinear form on g. 1. The root system R spans h∗ as a vector space. The elements hα , α ∈ R, defined by hα = 2Hα , (α, α) (123) span h as a vector space. 2. For any α ∈ R, the root subspace gα is one-dimensional. 3. For any α, β ∈ R, the number hhα , βi = 2(α, β) (α, α) (124) is an integer. 4. For any α, β ∈ R, the reflection operator sα : h∗ → h∗ , defined by sα (β) = β − hhα , βi α = β − 2(α, β) α (α, α) (125) is also a root. In particular, sα (α) = −α ∈ R. 5. For any α ∈ R, the only multiples of α that are also roots are ±α. 6. For any α, β ∈ R such that β 6= ±α, the subspace M V = gβ+kα (126) k∈Z is an sl(2, C)α -module. 7. For any α, β ∈ R such that α + β ∈ R, we have [gα , gβ ] = gα+β . We will not give the proof here. Rather we will try to understand the construction by studying the abstract root space Sec. 4.4.5. However, before leaving the Lie algebra and dive into the dual space, we derive the Cartan-Weyl basis which replies on taking the Killing form to be the metric (·, ·). 27 4.4.4 Cartan-Weyl basis In this section we derive the Cartan-Weyl basis. While the construction is not more insightful than what we have done in the last section, it is useful in physics by explicitly defining a usable basis for a semisimple Lie algebra. We would like to use the Hj ’s and Eα ’s defined in Eq. (105) as a basis of the Lie algebra g. To do this, it remains to uniquely define each eigenvector Eα and the corresponding root α given a basis {Hj } of h. With our construction this still requires two constraints: 1. a rescaling factor for the Hj and the corresponding root components αj , 2. a normalization condition for Eα . Both of these can be chosen by explicitly defining the bilinear form (the natural pairing in this case) on h and h∗ . Below we define such a natural pairing associated with the the Killing form K(a, b), a, b ∈ g. It is natural to consider the Killing form since we work in the adjoint rep. By Eq. (49), the Killing form is a symmetric bilinear invariant form. By Def. 0, the Killing form is nondegenerate given a semisimple Lie algebra g. Therefore we can identify it as a metric tensor on g, denoted as gab = K(a, b) ≡ (a, b). (127) Using the definition of the structure constant in Eq. (7) and the relation between the Killing form and the structure constants in Eq. (50), we can show the following results: 1. If α is a root, then −α is a root. 2. We can write the nondegenerate Killing form in a block diagonal. Denoting gjk = (Hj , Hk ) (128) gα,−α = (Eα , E−α ), (129) and we have gab gjk = 0 gα,−α gα,−α 0 0 gβ,−β gβ,−β 0 .. . , (130) where j, k ∈ {1, 2, · · · , l} and α, β label different roots. Notice here gjk is an l × l dimensional matrix. Nondegeneracy requires that we always have det [gjk ] 6= 0 and gα,−α = 6 0 for any root α. Both of these results have been mentioned in the previous section, but we show them here as a result of choosing the Killing form as the metric. The fact that gjk itself is a metric tensor on h allows us to use it as a prescription to uniquely define the bilinear form h·, ·i on h and h∗ . Below I give the standard procedure to construct the corresponding inner products in and between a vector space and its dual given a metric. This logic seems not intuitional to physicists. Consider an element in h as X Hλ = λ0j Hj , (131) j 28 where we write out the summation symbol on purpose, and λ0j ’s are some coefficients. With the dual space structure and the metric tensor, the element Hλ can also be expressed as Hλ = λj g jk Hk = λj Hj . (132) Here we have used the Einstein notation and asserted j g jk gkm = δm , (133) j where δm in matrix form is the identity matrix, and g jk is the inverse of the metric tensor gjk . We see that by defining λj = λ0j (134) we have created a bijective mapping between Hλ ∈ h and λ ∈ h∗ , defined uniquely by the metric tensor. We denote Eq. (132) by the bilinear form (inner product) hλ, Hi such that Hλ = hλ, Hi = λj g jk Hk = λj Hj . (135) Notice that Hλ is both the element in h corresponding to root λ but also the result of the bilinear form. The Killing form of two such elements Hα and Hβ in the Cartan subalgebra h is (Hα , Hβ ) = αj (Hj , Hk )β k = αj gjk β k = αj g jk βk . (136) For convenience, we denote the dual of the Killing form in h∗ also by the inner product αj g jk βk = (α, β). (137) Immediately, we see that our definition above satisfies (Hα , Hβ ) = (α, β) = hα, Hβ i , (138) which is consistent with Eq. (117). So far we have fixed the rescaling factor between Hα and α by connecting them with a metric. Now we determine the eigenvector Eα given its root α. Using the Jacobi identity (2) and Eq. (105), we can show [Hj , [Eα , Eβ ]] = (αj + βj )[Eα , Eβ ]. (139) 1. If αj + βj = 0 for some j, then [Eα , Eβ ] ∈ h, so α + β = 0. We then calculate ([Eα , E−α ], Hβ ) = (Eα , [E−α , Hβ ]) = hα, βi (Eα , E−α ) = hHα , Hβ i (Eα , E−α ), (140) where we have used the fact that the Killing form is invariant [this also proves Eq. (118)]. Thus we have [Eα , E−α ] = (Eα , E−α )Hα . (141) 2. If αj + βj 6= 0 for some j, then [Eα , Eβ ] ∈ g/h, so α + β 6= 0. Then [Eα , Eβ ] is an eigenvector of Hj with eigenvalue (αj + βj ). We write [Eα , Eβ ] = Nαβ Eα+β , where the matrix Nαβ is defined with the following properties: (a) The matrix Nαβ depends on the normalization of Eα ’s. (b) If α + β is not a root then Nα+β = 0. 29 (142) Choose the normalization (Eα , E−α ) = 1, we have the Cartan-Weyl basis. Definition 0 [Cartan-Weyl basis 31 ]: The Cartan-Weyl basis of a semisimple Lie algebra g is a basis of g that consists of {Hj } and {Eα }, where Hj ’s are a basis of the Cartan subalgebra h of g and Eα ’s are a basis of g/h. The defining Lie products are [Hj , Hk ] = 0, (143) [Hj , Eα ] = αj Eα , (144) [Eα , E−α ] = hα, Hi , (145) [Eα , Eβ ] = Nαβ Eα+β (α + β 6= 0). (146) Notice that although both semisimple Lie algebras over R [such as su(n)] and over C [such as sl(n, C)] have Cartan subalgebras, the Cartan-Weyl approach automatically introduces the complex extensions of the real Lie algebras since diagonalization generally results in complex eigenvalues and complex coefficients of basis vectors in the expression of eigenvectors by the fundamental theorem of algebra. In the adjoint rep, the latter means the complexification of real Lie algebras. As an example, consider su(2), where we normally choose basis {e1 , e2 , e3 } such that [e1 , e2 ] = e3 , [e2 , e3 ] = e1 , [e3 , e1 ] = e2 . (147) We choose {e3 } as the Cartan subalgebra without loss of generality and find its eigendecomposition in the adjoint rep, i.e., [e3 , e1 + ie2 ] = −i(e1 + ie2 ). (148) Therefore, the diagonalization of the Cartan subalgebra of the real su(2) algebra automatically leads to an expression that is only defined in its complex extension sl(n, C). This is also why we study sl(n, C) instead of su(n) in the first place. 4.4.5 Root system 4.4.5.1 Definition In Secs. 4.4.3–4.4.4 we see how the root system can tell us about the structure of a semisimple Lie algebra. In this section we examine the abstract root system as an independent object, which will help us understand Thm. 19. It can be proved that the dual space generated 32 by α ∈ R is completely determined by the real vector space generate by α ∈ R, denoted as h∗R , or in mathematical form h∗ = h∗R ⊕ ih∗R . Therefore below we only need to define the real vector space. Definition 0 [abstract root system]: An abstract root system is a finite set of elements R ⊂ E\{0}, where E is a Euclidean vector space 33 such that the following properties hold: 1. R generates E as a vector space. The number r = dim E is called the rank of R. 2. For any two roots α and β, the number nαβ = 2(α, β) (β, β) (149) is an integer. 31 The Cartan-Weyl basis is not unique. Using h , defined in Eq. (124), instead of H , we get the Chevalley basis where α α Nαβ become all integers. 32 Here a vector space “generated” by v means the minimum set of elements that contains v and has the structure required by a vector space (zero, linearity, closeness, etc.). Similarly for a Lie algebra generated by some element a. 33 A real vector space with an inner product. 30 3. Let sα : E → E be defined by sα (λ) = λ − 2(α, λ) α. (α, α) (150) Then for any roots α and β, sα (β) ∈ R. 4. If, in addition, R satisfies the property that if α and cα are both roots then c = ±1, then R is called a reduced root system. 34 Definition 0 defines the abstract root system as an independent mathematical construction, but from Thm. 19 we immediately realize the following theorem. Theorem 20 [semisimple Lie algebra and reduced root system]: Let g be a semisimple complex Lie algebra, with root decomposition given in Eq. (106). Then the set of roots R ⊂ h∗R \{0} is a reduced root system. Therefore we have turned the classification of all semisimple Lie algebras into the classification of all possible reduced root systems, although strictly speaking the one-to-one mapping is not yet established. To better understand the root system, we can give Def. 0 very specific geometric meanings. 1. The number nαβ is twice of the projection of α onto β, which is an integer. 2. The operator sα (λ) is the reflection of λ around the hyperplane Lα = {λ ∈ E | (α, λ) = 0}. (151) If (α, λ) = 0 then sα (λ) = λ. If λ = α then sα (α) = −α. We also define the coroot α∨ ∈ E ∗ such that hα∨ , λi = 2(α, λ) . (α, α) (152) Clearly, for the root system of of a semisimple Lie algebra, we have α∨ = hα . (153) Then from Eqs. (124)–(125) we have hα∨ , αi = 2, (154) nαβ = hα, β i , (155) ∨ sα (λ) = λ − hλ, α i α. ∨ (156) As an example, let {ej } be a basis of Rn , with the inner product being the dot product (ej , ek ) = δjk . Let n o X E = (λ1 , λ2 , · · · , λn ) ∈ Rn | λj = 0 (157) R = {ej − ek | 1 ≤ j, k ≤ n, j 6= k} ⊂ E. (158) and Then R is a reduced root system. The reflection sα is just the transposition of j, k entries for α = ej − ek , denoted as sjk . This root system is of rank n − 1, and historically called the root system of type An−1 . Clearly, this reduced root system is the root system of sl(n, C) given in Eq. (113). We next consider the possible geometric configurations a reduced root system can have. From conditions 2 and 3 of Def. 0, we realized that there is only a finite set of configurations (relative angles) two noncolinear roots α, β can have, which we will not shown here. For each configuration, its nαβ , nβα values can be determined correspondingly. A special class of automorphisms of a root system that keep nαβ invariant is the Weyl group. 34 In this work we only consider reduced root systems. 31 Definition 0 [Weyl group]: The Weyl group W of a root system R is the subgroup of GL(E) generated by reflections sα , α ∈ R. The Weyl group W is a finite subgroup in the orthogonal group O(E), and the root system R is invariant under the action of W . For any w ∈ W , α ∈ R, and we have sw(α) = wsα w−1 . Specifically, for R being the root system of type An−1 , W is the group generated by transpositions sjk , which is none other than the symmetric group Sn . This is why the Lie algebra sl(n, C) is closely connected to the symmetric group Sn : because the Weyl group of the reduced root system of sl(n, C) is Sn ! 4.4.5.2 Positive roots and simple roots It is possible to find for each root system some small set of “generating roots”. To do so requires some new concepts. Let t ∈ E be such that for any root α, (t, α) 6= 0. Then t is called a regular element of E. Then we decompose R as R = R+ t R− , with R+ = {α ∈ R | (α, t) > 0}, (159) and R− = {α ∈ R | (α, t) < 0}. Such a decomposition is called a polarization of R, which depends on our choice of t. The roots α ∈ R+ will be called positive roots, and the roots α ∈ R− will be called negative roots. Definition 0 [simple roots]: A positive root α ∈ R+ is a simple root if it can not be written as a sum of two positive roots. We denote the set of simple roots by Π ⊂ R+ . Since (α, t) can take only finitely many values, we can show that any positive root can be written as a sum of simple roots. Therefore we have the following theorem. Theorem 21 [simple roots are a basis]: Let R = R+ t R− ⊂ E be a root system. Then the simple roots form a basis of the vector space E. Another useful property of simple roots is that if α, β ∈ R+ are simple then (α, β) ≤ 0. From the geometry we can prove that every α ∈ R can be uniquely written as a linear combination of simple roots with integer coefficients, i.e., α= r X (160) nj αj j=1 where nj ∈ Z and {α1 , · · · , αr } = Π is the set of simple roots. For α ∈ R+ , all nj ≥ 0. For α ∈ R− , all nj ≤ 0. We define the height of a positive root α ∈ R+ by r r X X ht(α) = ht nj αj = nj ∈ N. (161) j=1 j=1 For the root system R of type An−1 , suppose we choose the polarization as R+ = {ej − ek | j < k}. (162) αj = ej − ej+1 , (163) Then the simple roots are with j ∈ {1, · · · , n − 1}. The height of any positive root is given by ht(ej − ek ) = k − j. 32 4.4.5.3 Root and weight lattices Due to the fact that every root can be written as a linear combination of simple roots with integer coefficients, it is natural to think about the roots as a root lattice. A lattice in a real vector space E is an abelian group generated by a basis in E. Even more importantly we can also construct a weight lattice for a root system. We make the following definitions. Definition 0 [root and weight lattices]: A root lattice is the abelian group Q ⊂ E generated by α ∈ R. A coroot lattice is the abelian group Q∨ ⊂ E ∗ generated by α∨ . A weight lattice P ⊂ E is the dual lattice of Q∨ , defined by P = {λ ∈ E | hλ, α∨ i ∈ Z, ∀α ∈ R} = {λ ∈ E | hλ, α∨ i ∈ Z, ∀α∨ ∈ Q∨ }. (164) Elements of P are called integral weights. Since we can identify the Cartan elements hα as the coroots α∨ , the elements of the weight lattice are indeed the generalization of the weights as seen in the sl(2, C) case in Sec. 4.4.1. This actually implies a weight decomposition for general semisimple Lie algebras, but we will discuss this later. Given a polarization of R and a set of simple roots Π, the root lattice is just M Q= Zαj (165) j for αj ∈ Π. The weight lattice is then P = λ ∈ E | λ, αj∨ ∈ Z, ∀αj ∈ Π . (166) Clearly, the root lattice is sublattice of the weight lattice, i.e., Q ⊂ P . We can define a basis of P by introducing the fundamental weights. Definition 0 [fundamental weights]: The fundamental weights ωj ∈ E are weights that satisfy hωj , αk∨ i = δjk . (167) Then we see that {ωj } is a basis of E and that P = M (168) Zωj . j Hence we have constructed two sets of basis of E, i.e., the simple roots and fundamental weights. 4.4.5.4 Weyl chambers We have seen that the choice of polarization depends not on the specific element t ∈ E but on the sign of (t, α). Therefore a polarization is invariant if we change t without crossing any hyperplane Lα . We call the connected component of the complement of the hyperplanes Lα as a Weyl chambers, denoted by C. In other words, a Weyl chamber C is a region of E bounded by hyperplanes Lα that uniquely defines a polarization by R+ = {α ∈ R | (α, t) > 0}, t ∈ C. (169) Conversely, given a polarization we can always define a corresponding positive Weyl chamber C+ by C+ = {λ ∈ E | (λ, α) > 0, ∀α ∈ R+ } = {λ ∈ E | (λ, αj ) > 0, ∀αj ∈ Π}. (170) Therefore there is a bijection between all polarizations and all Weyl chambers. For different Weyl chambers, the Weyl group acts transitively on the set of Weyl chambers, meaning that Cl = sβl (Cl−1 ) 33 (171) for two adjacent Weyl chambers separated by the hyperplane Lβl . Given a polarization (Weyl chamber) and a set of simple roots Π, it is then straightforward to see that we can recover the complete root space from simple reflections defined be Weyl group elements sj ≡ sαj acting on simple roots αj ’s. For the root system An−1 , the Weyl group is Sn , and the simple reflections sj ’s are transpositions of the jth and j + 1th entries. Then we can define the positive Weyl chamber to be C+ = {(λ1 , · · · , λn ) ∈ E | λ1 ≥ · · · ≥ λn }. All the other Weyl chambers are obtained by applying to C+ permutations σ ∈ Sn , i.e., Cσ = (λ1 , · · · , λn ) ∈ E | λσ(1) ≥ · · · ≥ λσ(n) . (172) (173) We will discuss the relation to Sn in Sec. ??. 4.4.5.5 Dynkin diagrams We have seen that given a set of simple roots we can recover a root system. Therefore classifying root systems is equivalent to classifying possible sets of simple roots. We first define reducible and irreducible root systems. Definition 0 [reducibility of root systems]: A root system R is reducible if it can be written in the form R = R1 t R2 with R1 ⊥ R2 . Otherwise R is called irreducible. It can be shown that every reducible root system can be uniquely written in the form R1 t · · · Rn where Rj ’s are mutually orthogonal irreducible root systems. Therefore it suffices to classify irreducible root systems. Suppose R is an irreducible reduced root system and we have chosen a set of simple roots Π. To uniquely classify the simple roots, we define the Cartan matrices. Definition 0 [Cartan matrices]: The Cartan matrix A of a set of simple roots Π ⊂ R is the r × r matrix with entries ajk = nαj αk = αj∨ , αk = 2(αj , αk ) . (αk , αk ) (174) Immediately we have the following properties of the Cartan matrix: 1. For any j, ajj = 2. 2. For any j 6= k, ajk is a nonpositive integer. 3. For any j 6= k, ajk akj = 4 cos2 ϕ, where ϕ is the angle between αj and αk . If ϕ 6= π/2, then |αj | 2 |αk | 2 = akj . ajk For An−1 , r = n − 1, and the Cartan matrix is of dimension (n − 1) × (n − 1) that has the form 2 −1 −1 2 −1 −1 2 −1 . A= .. .. .. . . . −1 2 −1 −1 2 (175) (176) The information contained in the Cartan matrix can be presented in a graphical way, called the Dynkin diagram. 35 35 There are other kinds of diagrams in the study of semisimple Lie algebras. For example, the Satake diagrams classify simple Lie algebras over R. 34 Definition 0 [Dynkin diagram]: Let Π be a set of simple roots of a root system R. The Dynkin diagram of Π is the graph constructed in the following manner. 1. For each simple root αj , we construct a vertex vj of the Dynkin diagram. 2. For each pair of simple roots αj 6= αk , we connect the corresponding vertices by n edges, where n depends on the angel ϕ between αj and αk : π , n = 0. 2 2π , n = 1. (b) For ϕ = 3 3π (c) For ϕ = , n = 2. 4 5π , n = 3. (d) For ϕ = 6 (a) For ϕ = 3. For every pair of distinct simple roots αj 6= αk , if |αj | 6= |αk | and they are not orthogonal, we orient the corresponding (multiple) edge by putting on it an arrow pointing from the longer root to the shorter root. The Dynkin diagram has the following properties. 1. The Dynkin diagram of Π is connected iff R is irreducible. 2. The Dynkin diagram determines the Cartan matrix A. 3. R is determined by the Dynkin diagram uniquely up to an isomorphism. Therefore the classification of all irreducible root systems reduces to finding all the possible Dynkin diagrams of irreducible root systems. Theorem 22 [classification of Dynkin diagrams]: Let R be a reduced irreducible root system. Then its Dynkin diagram is isomorphic to one of the diagrams below 36 . • An (n ≥ 1): • Bn (n ≥ 2): • Cn (n ≥ 2): • Dn (n ≥ 4): • E6 : • E7 : • E8 : • F4 : 36 In each diagram the subscript is equal to the number of vertices. 35 • G2 : In our work, we only focus on the simply-laced An type 37 , which corresponds to the sl(n, C) Lie algebra. For completeness, below we give the special cases when n is smaller than required in the diagrams above. 1. For n = 1, A1 = B1 = C1 , which corresponds to the Lie algebra isomorphisms sl(2, C) ' so(3, C) ' sp(1, C). 2. For n = 2, B2 = C2 , which corresponds to the Lie algebra isomorphism so(5, C) ' sp(2, C). 3. For n = 2, D2 = A1 ∪ A1 , which corresponds to so(4, C) ' sl(2, C) ⊕ sl(2, C). 4. For n = 3, D3 = A3 ∪ A3 , which corresponds to so(6, C) ' sl(4, C). Other than these special cases, all root systems listed in Thm. 22 are distinct. 4.4.6 Classification of semisimple Lie algebras With the discussion on root systems classification of Dynkin diagrams, we can now go back to the classification of semisimple Lie algebras. We have shown that every semisimple Lie algebra defines a reduced root system. In fact, we can also recover a semisimple Lie algebra from a reduced root system. This is shown by the following theorems. We first have the triangular decomposition of semisimple Lie algebras. Theorem 23 [triangular decomposition]: Let g be a semisimple Lie algebra with root system R ∈ h∗ , and (·, ·) be a nondegenerate invariant symmetric bilinear form on g. Let R = R+ t R− be a polarization of R and Π = {α1 , · · · , αr } be the corresponding system of simple roots. Then the subspaces M n± = gα (177) α∈R± are subalgebras in g, and g = n− ⊕ h ⊕ n+ . (178) This comes directly from the root decomposition theorem and the fact that the sum of positive roots is positive. Given the triangular decomposition, we have the Serre relations 38 . Theorem 24 [Serre relations]: Given the definitions in Thm. 23, let ej ∈ gαj , fj ∈ g−αj be chosen such that 2 (ej , fj ) = , (179) (αj , αj ) and let hj = hαj ∈ h be defined by Eq. (123), i.e., hj = 2Hαj . (αj , αj ) (180) Then for j ∈ {1, · · · , r}, {ej } generate n+ , {fj } generate n− , and {hj } form a basis of h. In particular {ej , fj , hj } generate g. Moreover, the elements ej , fj , hj satisfy the following Serre relations: [hj , hk ] = 0, (181) [hj , ej ] = ajk ek , (182) [hj , fj ] = −ajk ek , (183) [ej , fk ] = δjk hj , (184) 1−ajk ek = 0, (185) 1−ajk fk = 0, (186) [ad(ej )] [ad(fj )] 37 A Dynkin diagram with no multiple edges is called a simply-laced diagram (ADE types). Chevalley-Serre relations. 38 Or 36 where ajk = nαj αk = αj∨ , αk (187) are the entries of the Cartan matrix. Each Cartan matrix ajk determines a unique semisimple complex Lie algebra via the Serre relations. In fact, one can show the following theorem. Theorem 25 [bijection between root systems and semisimple Lie algebras]: There is a natural bijection between the set of isomorphism classes of reduced root systems and the set of isomorphism classes of finitedimensional complex semisimple Lie algebras. Moreover, the Lie algebra is simple iff the reduced root system is irreducible. Combining Thms. 22 and 25, we have the following result. Theorem 26 [classification of simple Lie algebras]: Simple finite-dimensional complex Lie algebras are classified by Dynkin diagrams An , Bn , Cn , Dn , E6 , E7 , E8 , F4 , and G2 listed in Thm. 22. 4.4.7 Weight decomposition We can now classify the reps of complex semisimple Lie algebras, which reduces to the classification of irreps and finding a way to determine the multiplicities. To do this, we consider the weight decomposition of a finite-dimensional rep. Definition 0 [weight, general]: Let V be a g-module. A vector v ∈ V is called a vector of weight λ ∈ h∗ if for any h ∈ h we have hv = hλ, hi v. (188) The space of all vectors of weight λ is called the weight space and denoted as V [λ], i.e., V [λ] = {v ∈ V | hv = hλ, hi v, ∀h ∈ h}. (189) If V [λ] 6= {0} then λ is called a weight of V . Definition 0 is the generalization of Def. 0 for the sl(2, C) case. The set of all weights of V is denoted by P (V ), i.e., P (V ) = {λ ∈ h∗ | V [λ] 6= 0}. (190) Since vectors of different weights are linearly independent, P (V ) is finite for finite-dimensional reps. Theorem 27 [weight decomposition, general]: Every finite-dimensional rep of g admits a weight decomposition given by M V = V [λ]. (191) λ∈P (V ) Moreover, all weights of V are integral, meaning that P (V ) ⊂ P , where P is the weight lattice defined in Def. 0. This comes from Thms. 17 and 19. As in the sl(2, C) case this weight decomposition agrees with the root decomposition of g, meaning that if x ∈ gα , then x.V [λ] ⊂ V [λ + α]. We next study the dimensions of the weight subspaces V [λ]. This is done by defining the formal generating series for these dimensions, i.e., the character of V . 37 Definition 0 [character, weight decomposition]: Let V be a finite-dimensional g-module. Then its character ch(V ) is defined by X ch(V ) = (dimV [λ])eλ . (192) λ for λ ∈ P (V ). Here ch(V ) ∈ C[P ] where C[P ] is the algebra generated by formal expressions eλ , λ ∈ P , subject to relations eλ eµ = eλ+µ and e0 = 1. It can be shown that algebra C[P ] is isomorphic to the algebra of Laurent polynomials in r variables, where r is the rank of g. Therefore, we can understand the character as a series α of polynomials. As an example, for sl(2, C), we have P = Z . So C[P ] is generated by enα/2 , n ∈ Z. 39 2 Denoting eα/2 = x, we have C[P ] = C[x, x−1 ]. Thus by Thm. 17 the character of each irreducible Vn is given by ch(Vn ) = xn + xn−2 + · · · + x−n = xn+1 − x−(n+1) . x − x−1 (193) We have already seen the definition of characters for real compact Lie groups and Lie algebras in Sec. 4.3.5. Although both characters can be used to calculate the multiplicities of irreps, in Sec. 4.3.5 the characters are calculated from a group approach while here the characters are calculated using the weight decomposition of Lie algebras. 40 The characters have the following properties. 1. ch(C) = 1. 2. ch(V1 ⊕ V2 ) = ch(V1 ) + ch(V2 ). 3. ch(V1 ⊗ V2 ) = ch(V1 )ch(V2 ). 4. ch(V ∗ ) = ch(V ), where · is defined by eλ = e−λ . In the sl(2, C) case we see that the characters are symmetric with respect to Weyl group actions (x → x−1 ) in this case. More generally, we have the following theorem. Theorem 28 [dimension of weight spaces are Weyl invariant]: If V is a finite-dimensional g-module, then the set of weights and dimensions of weight subspaces are Weyl group invariant, meaning that for any w ∈ W , dim V [λ] = dim V [w(λ)]. (194) w.ch(V ) = ch(V ), (195) Equivalently, we can write where the action of W on C[P ] is defined by w.eλ = ew(λ) . We now consider irreps and their properties. Definition 0 [highest weight rep]: A nonzero rep ρV of g is called a highest weight rep if V (not necessarily finite-dimensional) is generated by a vector v ∈ V [λ] such that x.v = 0 for all x ∈ n+ , where n+ is defined by the triangular decomposition of g in Thm. 23. Then v is called the highest weight vector, and λ is the highest weight of V . The highest weight rep is important because of the following theorem. Theorem 29 [every finite-dimensional irrep is highest weight rep]: Every finite-dimensional irrep of g is a highest weight rep. 39 ??? 40 I am not perfectly clear of the differences and relations between these two definitions. It seems like they are self-consistent. Anyway I will leave some questions marks here??? 38 In any highest weight rep there is a unique highest weight and unique-up-to-a-scalar highest weight vector. In any highest weight rep with highest weight vector vλ ∈ V [λ], the following conditions hold: hvλ = hh, λi vλ , xvλ = 0, ∀h ∈ h, ∀x ∈ n+ . (196) (197) We can define a module Mλ as the g-module generated by a vector vλ satisfying only Eqs. (196)–(197). Formally, M λ is called the Verma module, but we will not dig into the details at this time. The point of introducing the Verma module is... here put comments from youtube video By Thm. 29, to classify all finite-dimensional irreps is to classify all highest weight reps which are finitedimensional and irreducible. Theorem 30 [existence of an irreducible highest weight rep]: For any λ ∈ h∗ , there exists a unique up to isomorphism irreducible highest weight rep with highest weight λ. This g-module of this rep is denoted Lλ . For sl(2, C), if λ ∈ N, then Lλ = Vλ is the finite-dimensional irreducible module of dimension λ + 1. If λ∈ / N, then Lλ = Mλ . More generally, for any Lie algebra, for “generic” λ, the Verma module is irreducible so Mλ = Lλ . Also by Thm. 29, every irreducible finite-dimensional V must be isomorphic to one of Lλ . Thus to classify all finite-dimensional irreps of g, we need to find out which of Lλ are finite-dimensional. We make the following definitions. Definition 0 [dominant weight]: A weight λ ∈ h∗ is called a dominant integral weight or a dominant weight if for all α ∈ R+ , hλ, α∨ i ∈ N. (198) The set of all dominant integral weights is denoted P+ . The point of introducing dominant integral weights is the following theorem. Theorem 31: For every λ ∈ P+ , Lλ is irreducible and finite-dimensional. These reps are pairwise nonisomorphic, and every irreducible finite-dimensional rep is isomorphic to one of them. To study the structure of the reps, one requires the BGG resolution and Weyl character formula, which are beyond our scope. At this point, we know we can write any g-module as M V = nλ Lλ . (199) λ∈P+ The multiplicities nλ can be found by writing the character ch(V ) in the basis ch(Lλ ). We skip the derivations and only give the method to calculate multiplicities for sl(n, C) in the next section. 4.4.8 Reps of sl(n, C) In this section, we classify all of the irreducible reps of sl(n, C) both as an example and as our purpose in this work. 41 Recall the root system of sl(n, C) is given by R = {ej − ek | j 6= k} ⊂ h∗ = Cn /C(1, · · · , 1). (200) The set of positive roots is given by R+ = {ej − ek | j < k}. (201) 41 I do not know if in my life I will need to use any of the knowledge above on any other Lie algebras, but at least knowing these stuff is fun. 39 The weight lattice and the set of dominant roots are P = {(λ1 , · · · , λn ) ∈ h∗ | λj − λk ∈ Z}, (202) P+ = {(λ1 , · · · , λn ) ∈ h | λj − λj+1 ∈ N}. (203) ∗ Since adding a multiple of (1, · · · , 1) does not change the weight lattice and the dominant roots (changing P the basis of h∗ by moving along the line ej = 0), we can represent P and P+ as P = {(λ1 , · · · , λn−1 , 0) | λj ∈ Z}, (204) P+ = {(λ1 , · · · , λn−1 , 0) | λj ∈ N, λ1 ≥ · · · ≥ λn−1 ≥ 0}. (205) Equation (205) means that the set of dominant integer weights for sl(n, C) can be identified with the set of partitions with n − 1 parts, which can then be represented graphically by the Young diagrams. We will introduce the Young diagrams and Young tableaux in Sec. ??. eg 8.42 eg8.43 eg 8.44 ex8.4 ex8.5 ex8.7 ex8.9 P Any irreducible finite-dimensional Lλ as an sl(n, C)-module appears as a subspace in (Cn )N , N = λj , determined by suitable symmetry properties, i.e. transforming ina certain way under the action of symmetric group Sn . We can calculate the characters of irreps of sl(n, C). Theorem 32 [Weyl character formula, sl(n, C)]: Let λ = (λ1 , · · · , λn ) ∈ P+ be a dominant weight for sl(n, C), λj ∈ N, λ1 ≥ λ2 ≥ · · · ≥ λn . Then the character of the corresponding irrep of sl(n, C) is given by ch(Lλ ) = Aλ1 +n−1,λ2 +n−2,··· ,λn Aλ1 +n−1,λ2 +n−2,··· ,λn Q = , An−1,n−2,··· ,0 j<k (xj − xk ) (206) where Aµ1 ,µ2 ,··· ,µn = det xµj k 1≤j,k≤n = X 1 n sgn(s)xµs(1) · · · xµs(n) . (207) s∈Sn One may recognize that in Eq. (206), the nominator is the general alternating polynomial, the denominator is the Vandermonde polynomial, and the RHS itself is the Schur polynomial. 4.4.9 Casimir operators The adjoint representation is irreducible iff the Lie algebra is simple, since subrepresentations correspond to ideals. have I included this??? casimir elements 6.15 for quadratic, others? given in Greiner p517 Can find casimir operaters this way greiner 115, 137 greiner chap 12 pf 112-117 simplicity pf p6 lilian notes 40 For a Lie algebra, casimir operators are defined such that they commute with every generator of the Lie algebra, and can therefore be used to label irreducible representations in the Cartan-Weyl basis42 . We can now define the rank of a Lie group. There are no general method to get all the Casimir operators for an arbitrary Lie algebra. However, for SU(n), there are methods available43 . In general there is no method to construct the casimir operators for arbitrary semisimple groups. Casimir operators are not unique. quadratic casimir operators are conserved quantity for 2 dim phase space Miller 50 Racah theorem: Greiner 109 Each multiplet of semisimple Lie group can be uniquely characteried by the eigenvalues C1 C2 Cl of the l Casimir operators C1, C2, Cl One of the Casimir operators is always given by the quadratic form Griener 116 Harish-Chandra isomorphism, 8.8 Kirillov 4.5 4.5.1 Direct product to direct sum Symmetric and antisymmetric tensors bilinear form decomposition into symmetric and anti-symmetric invariant bilinear form in rep theory; G-invariant Define the averaging map Av : V ⊗n → V ⊗n such that n n O 1 X O Av( vj ) = vσ(1) n! j=1 j=1 (208) σ∈Sn Then the symmetric map is defined as Symn V = Image(Av : V ⊗n → V ⊗n ). If V is a rep then 1. Av is a morphism of reps V ⊗n → V ⊗n (interwining map) 2. Image of a morphism is a subrep. Symn V is a homogeneous polynomials of degree n in a basis of V . Symn C2 is (n + 1)-dimensional. Symn C2 : su(2) → gl(n + 1, C). This gives the state multiplet. Symmetric tensors and skey-symmetric/antisymmetric tensors Λ2 C2 the exterior square, antisymmetric tensors Sym2 Sym2 C2 ' Sym4 C2 ⊕ C. This looks like quadratic forms. invariant theory 4.5.2 Direct product into direct sum Direct sum is reducible. Direct product may be irreducible. Symmetric powers may be irreducible. But symmetric powers have a subrep that is irrep, the symmetric tensors. Sym2 C2 = 12 {e1 ⊗ e1 , e2 ⊗ e2 , e1 ⊗ e2 + e2 ⊗ e1 } direct product and sum of semisimple Lie algebras are both semisimple Lie algebras 4.5.3 The Clebsch-Gordan coefficients (CG coefficients) For physicists, one of the most useful properties theorem of complete reducibility. 44 42 I regarding reducibility and simplicity is the following will add more to the Cartan-Weyl basis, roots, and multiplets. Cvitanović in Reference 44 One can also look at the concept of reductive algebras. The su(n) Lie algebra is semisimple and reductive. The rep of su(n) is reducible and decomposable. 43 See 41 Theorem 33 [Weyl’s theorem on complete reducibility 45 ]: Every finite-dimensional rep of a semisimple algebra decomposes into the direct sum of irreducible reps (or irreps). One should also note the difference between reducibility and decomposability. Without proof, we state here that a rep is reducible if under similarity transformation it can be written in a nontrivial upper triangular block form (Lie’s theorem). A rep is decomposable if under similarity transformation it can be written in a nontrivial block diagonal form. Decomposability implies reducibility, but not conversely. We can define the direct product of two reps. For example, the direct product of su(n) reps is used across quantum physics. By Thm. 33, The CG coefficients are used to perform the explicit direct sum decomposition of the tensor product of irreducible representations in cases where the numbers and types of irreducible components are already known abstractly. The detailed constructions and derivations will be shown as an example for su(2) in Sec. 5.3.4. The CG coefficients for a general Lie algebra is unknown. However, the algorithms to produce CG coefficients for su(n) are known. weights are connected here Summary of properties of sl(n, C) and su(n) 4.6 1. Lie algebra. Let g = sl(n, C) and h = {diag{c1 , c2 , · · · , cn } | cj ∈ C, P cj = 1}. Define ej ∈ h as the functional ∗ h1 ej : .. → hj . . (209) hn Then h∗ = M Cei /C X ej (210) and E = h∗R = M Rej /R X ej . (211) The inner product is defined by (λ, µ) = λj µj if λ and µ are chosen such that P λj = P (212) µj = 0. 2. Root system. The root system is R = {ej − ek | j 6= k}. (213) The root subspace corresponding to α = ej − ek is gα = CEjk (214) hα = Ejj − Ekk . (215) and the corresponding coroot hα = α∨ ∈ h is 45 See Kirillov and Fulton for Weyl’s construction with radicals and nilpotents. 42 3. Positive and simple roots. We choose the set of positive roots as R+ = {ej − ek | j < k} (216) with |R+ | = n(n − 1) . 2 (217) The set of simple roots is Π = {α1 , · · · , αn−1 }, (218) αj = ej − ej+1 . (219) 4. Dynkin diagram. 5. Cartan matrix. 2 −1 A= −1 2 −1 −1 2 .. . −1 .. . −1 .. . 2 −1 . −1 2 (220) 6. Weyl group. The Weyl group is W = Sn (221) sj =?? (222) acting on E by permutations. Simple reflections are 7. Weight and root lattices. The weight lattice is P = {(λ1 , · · · , λn ) | λj − λk ∈ Z}/R(1, · · · , 1) = {(λ1 , · · · , λn−1 , 0) | λj ∈ Z}. (223) n o X Q = (λ1 , · · · , λn ) | λj ∈ Z, λj = 0 . (224) ∼ Z/nZ. P/Q = (225) The root lattice is 43 8. Dominant weights and Weyl chamber. C+ = {(λ1 , · · · , λn ) | λ1 > · · · > λn }/R(1, · · · , 1) = {(λ1 , · · · , λn−1 , 0) | λ1 > · · · > λn−1 > 0}. (226) P+ = {(λ1 , · · · , λn ) | λj − λj+1 ∈ N}/R(1, · · · , 1) = {(λ1 , · · · , λn−1 , 0) | λj ∈ Z, λ1 ≥ · · · ≥ λn−1 ≥ 0}. (227) As algebras, Then for simple Lie algebras, the kernel is {0}, which means that all the nontrivial reps of simple Lie algebras are faithful.We can prove that the kernel of a rep of Lie algebra g is the ideal of g. Reps of su(n). Here we list the properties of su(n) regarding reducibility and multiplets: 1. Since sl(n, C) is semisimple, any complex finite-dimensional rep of su(n) is completely reducible. 2. All defining reps of su(n) are irreps. 3. All nontrivial reps of su(n) are faithful. 4. For each m ∈ {2, 3, · · · }, su(n) has exactly one irrep on Cm up to isomorphism. 5. The fundamental multiplet of su(n) is the same as the vector space associated to its natural/defining/standard rep, both being Cn . character of a rep??? The su(2) Lie Algebra and its reps 5 In the next three sections we discuss in detail the su(2), su(3), and su(4) Lie algebra. 5.1 Generators and structure constants By Eq. (8), the su(2) Lie algebra is of dimension 3. From Eqs. (12)–(13), the 3 generators of su(2) are given by the Pauli matrices, i.e., 0 1 0 −i 1 0 x y z λ1 = σ = , λ2 = σ = , λ3 = σ = . (228) 1 0 i 0 0 −1 The corresponding basis, according to Eq. (15), are i x i 0 1 i y i 0 e1 = − σ = − , e2 = − σ = − 2 2 1 0 2 2 i −i , 0 i z i 1 e3 = − σ = − 2 2 0 0 . −1 (229) The structure constants are found by Eq. (19). There is only one nonzero independent structure constant up to the completely antisymmetric nature, i.e., f123 = 1 Tr [[σ x , σ y ]σ z ] = 1. 4i This agrees with our knowledge of the Pauli matrices. 44 (230) 5.2 5.2.1 The Lie correspondence between SU(2) and su(2) From SU(2) to su(2) Let U ∈ SU(2) be a 2 × 2 unitary matrix with unity determinant. One way to parameterize U with real parameters xj , j ∈ {1, 2, 3}, is q − 1 − 14 xj xj − 2i x3 − 21 x2 − 2i x1 . q U = (231) 1 i 1 i x − x − 1 − x x + x 2 2 2 1 4 j j 2 3 Using Eq. (30), we can check that this choice of parameterization matches our basis choice in Eq. (229), i.e., (el )jk = 5.2.2 ∂ Ujk (x1 , x2 , x3 )|x1 =x2 =x3 =0 . ∂xl (232) From su(2) to SU(2) By Eq. (31), given a basis in Eq. (229), any matrix U ∈ SU(2) can be parameterized by i U (x1 , x2 , x3 ) = exp (xj ej ) = exp − (x1 σ x + x2 σ y + x3 σ z ) . 2 (233) One can check for self-consistency by combining Eqs. (232) and (233). 5.3 5.3.1 Reps Defining rep By Def. 0, the su(2) Lie algebra over R constitutes all the anti-Hermitian 2 × 2 matrices with vanishing n o traces. Thus the defining rep is none other than the 2-dimensional rep that we label as (e1 )jk , (e2 )jk , (e3 )jk , where the basis operator matrices {ej } are given in Eq. (229). The vector space of the rep is C2 . 5.3.2 Adjoint rep The adjoint rep of su(2) is of dimension 3. The adjoint matrices of the three basis elements are found by Eq. (41), i.e., k=1 j=1 ad(e1 )jk = f1kj = j=2 j=3 k=2 0 0 0 0 0 1 k=3 0 −1 , 0 0 0 1 = 0 0 0, −1 0 0 0 −1 0 = 1 0 0, 0 0 0 (234) ad(e2 )jk = f2kj ad(e3 )jk = f3kj (235) (236) where n we have detailed the labels o for ad(e1 ). Thus the adjoint rep is the 3-dimensional rep ad(e1 )jk , ad(e2 )jk , ad(e3 )jk . The vector space of the rep is the su(2) Lie algebra itself. The Killing form of su(2) can be found following Eq. (51), i.e., −2 0 0 gjk = K(ej , ek ) = 0 −2 0 = −213×3 . 0 0 −2 45 (237) 5.3.3 Irrep fast finish!!!! this is isomorphism !!! Kirillov p47 Theorem 34 [classification of su(2)]: Any irrep of su(2) is isomorphic to Symn C2 for some n ∈ N. has weight diagram −n −(n − 2) n − 2 n 46 This (238) For any n ∈ N+ there is one and only one irrep of dimension n for su(2) up to isomorphism. 1 H= 0 0 −1 0 E= 0 1 0 0 F = 1 0 . 0 (239) [H, E] = 2E, (240) [H, F ] = −2F, (241) [E, F ] = H. (242) The su(2) Lie algebra has irreps. Consider basis operators specifying the dimension of h {eµi } without h i h i 1 (j) (j) (j) e1 , e2 , e3 , for j ∈ N = its state space. A (2j + 1)-dimensional irrep is found by 2 µν µν µν 1 3 47 0, , 1, , · · · , where the basis operators are defined as 2 2 e1 = −iJ x , (243) e2 = −iJ , (244) e3 = −iJ . (245) y z Here physicists will recognize the J µ operators as the angular momentum operators. With the definitions J ± = J x ± iJ y , 2 (246) 2 2 2 J 2 = (J x ) + (J y ) + (J z ) = (J z ) + J z + J − J + , the angular momentum operators satisfy (247) 48 J 2 |jmi = j(j + 1) |jmi , (248) J |jmi = m |jmi , p J ± |jmi = (j ∓ m)(j ± m + 1) |j(m ± 1)i , (249) z (250) where we have defined the orthomormal basis states with the azimuthal quantum number j and the magnetic 1 quantum number m such that given j ∈ N, m ∈ {j, j − 1, · · · , −j + 1, −j}. 2 Here are some properties of the irreps of su(2): 1. The fundamental multiplet. 1 The fundamental multiplet is found when 2j + 1 = 2, i.e., j = . We can write the fundamental 2 1 1 1 1 multiplet as the state multiplet , , ,− . When the other j values are irrelevant, the 2 2 2 2 n = 0, Sym0 C2 = C case j = 0 is the trivial rep. 48 Derivations of these relations can be found in any quantum mechanics textbook. 46 For 47 The 46 fundamental multiplet can also be labelled as {|1i , |0i}, {|↑i , |↓i}, or {|ei , |gi} within different contexts. This agrees with our discussion in Sec. 4.3.4 that the fundamental multiplet for su(2) is C2 . The fact that the state space of the defining rep is also C2 means that in physics, we define su(2) with its fundamental rep conventionally. 1 For j = , we have 2 1 1 i 1 x ( ) x (2) e1 2 = − σ x = σ (J ) 1 2 2 J ± ( 2 ) = σ± 1 1 1 y i (2) y (2) (251) , = σ , (J ) e2 = − σ y , 1 3 ) ( 2 2 2 2 = 1 J 2×2 1 (2) 1 4 i (J z )( 2 ) = 1 σ z e = − σz 3 2 2 which agrees with Eq. (229). 2. Higher-dimensional multiplets and explicit operator forms. 1 For arbitrary j ∈ N, the corresponding (2j+1)-dimensional multiplet is the state multiplet {|j, m = ji , |j, m = j − 1i , · 2 For j = 1, we have √ 2 0 0 √ √ √ 1 (J x )(1) = 2 0 0 2 0 2 √ √ (1) 2 J+ = 0 0 2 0 2 0 (1) (1) √ 0 0 0 e1 = −i(J x ) 0 0 − 2 √ i √ (1) (1) (1) 0 0 0 , . (252) (J y ) = 2 0 − 2 , e2 = −i(J y ) √ √ (1) 2 = 2 0 0 J− (1) (1) 2 0 0 z √ e3 = −i(J ) 2 0 0 1 0 0 (1) 2 (1) (J z ) = 0 0 0 J = 213×3 0 0 −1 Notice that the multiplet with j = 1 (the triplet rep) is of the same dimension as the state space of the adjoint rep. It turns out that the adjoint rep of su(2) given in Eqs. (234)–(236) is an equivalent rep of the triplet rep given in Eq. (252). In fact, su(2) has a unique (2j + 1)-dimensional irrep up to equivalent classes. 3. The sl(2, C) Lie algebra. We notice that {J + , J − , J z } over C is also a Lie algebra given the commutations + − J , J = 2J z , , J −, J z = J −, , [J z , J x ] = J + . (253) In fact, the Lie algebra {J + , J − , J z } over C is the complexified Lie algebra [i.e., sl(2, C)] of the Lie algebra {e1 , e2 , e3 } [i.e., su(2)]. Their transformations are given by i + e1 − 2 − 2i J 0 1 e2 = − 1 J − , 0 2 2 e3 0 0 −i J z + i −1 0 e1 J J − = i 1 0e2 . Jz 0 0 i e3 47 (254) (255) The transformations between the generators {λ1 , λ2 , λ3 } and {J + , J − , J z } are 49 λ1 1 1 0 J+ λ2 = −i i 0J − , λ3 0 0 2 Jz + 1 i J 0 λ1 2 2 J − = 1 − i 0λ2 . 2 2 Jz 0 0 i λ3 (256) (257) The sl(2, C) and su(2) have the same set of multiplets. 4. Notation. For future use, we use the notation [j] to represent the (2j + 1)-multiplet. 5.3.4 CG coefficients We introduce the decomposition of direct products of sl(2, C) irreps in this section. Since sl(2, C) and su(2) have the same set of multiplets, the direct sum of multiplets of sl(2, C) resulted from the decomposition is also that of su(2). 5.3.4.1 Direct product of reps Consider the direct product of two sl(2, C) irreps J1± , J1z and J1± , J1z with state multiplets [j1 ] ≡ {|j1 m1 i} and [j2 ] ≡ {|j2 m2 i}. Without loss of generality, we let j1 ≥ j2 . The resulting rep has operators that act on the vector space given by [j1 ] ⊗ [j2 ] ≡ {|j1 m1 i ⊗ |j2 m2 i}, where the basis |j1 m1 i ⊗ |j2 m2 i ≡ |j1 m1 , j2 m2 i is named the product basis. We introduce the total operators J µ = Jµ ⊗ 1 + 1 ⊗ Jµ (258) for µ ∈ {x, y, z, +, −}. Then it is straightforward to check that {J x , J y , J z } is the su(2) Lie algebra and {J ± , J z } is the sl(2, C) algebra. It can be shown that the rep of {J ± , J z } acting on the vector space [j1 ]⊗[j2 ] is reducible 50 . The state basis in which the rep is in block diagonal form (manifestly reducible) is called the total basis labelled as [J] ≡ |JM i, where J ∈ {j1 + j2 , j1 + j2 − 1, · · · , j1 − j2 }, M ∈ {J, J − 1, · · · , −J}. Indeed, the dimensions of the vector spaces {|JM i} and {|j1 m1 , j2 m2 i} are both (2j1 + 1)(2j2 + 1). 5.3.4.2 CG coefficients Completeness of basis states allows us to write X |JM i = |j1 m1 , j2 m2 i hj1 m1 , j2 m2 |JM i . (259) m1 ,m2 Here, the coefficient hj1 m1 , j2 m2 |JM i is the CG coefficient we are looking for. It has the following properties: 1. Triangle inequality. hj1 m1 , j2 m2 |JM i = 6 0 only if j1 − j2 ≤ J ≤ j1 + j2 . 2. hj1 m1 , j2 m2 |JM i = 6 0 only if m1 + m2 = m. 3. By the Condon-Shortley convention, (a) The CG coefficients are chosen to be real. (b) hj1 j1 , j2 (J − j1 )|JJi is positive. 49 This 50 See is consistent with the choice in my thesis which is nice. Chap. 3.5, Pfeifer. 48 4. hj1 m1 , j2 m2 |JM i = (−1) j1 +j2 −J hj1 (−m1 ), j2 (−m2 )|J(−M )i. 5. Since CG coefficients relate one orthonormal basis to another, the CG matrix is unitary. In the Condon-Shortley convention, it is orthogonal, i.e., hj1 m1 , j2 m2 |JM i = hJM |j1 m1 , j2 m2 i. 6. The algorithm to find CG coefficients: (a) Start from |j1 + j2 , j1 + j2 i = |j1 j1 , j2 j2 i. (b) Use the lowering operator to get all |j1 + j2 , M i for M ≥ 0. (c) Find |j1 + j2 − 1, j1 + j2 − 1i using properties 1–3. (d) Using the lowering operators to get all |j1 + j2 − 1, M i for M ≥ 0. (e) Loop until the end. (f) Using property 4 to get the M < 0 entries. Given the CG coefficients, we can transform the reps of {J ± , J z } between the product basis and the total basis by X hJM 0 |J µ |JM i = hJM 0 |j1 m01 , j2 m02 i hj1 m01 , j2 m02 |J µ |j1 m1 , j2 m2 i hj1 m1 , j2 m2 |JM i . (260) m1 m2 m01 m02 for µ ∈ {±, z}. By definition the J µ operators are in block diagonal forms in the total basis. Other notations of the CG coefficients in the literature include √ j1 j2 J jj 0 j1 −j2 +M 2J + 1 . hj1 m1 , j2 m2 |JM i = Smm0 ,JM = (−1) m1 m2 −M 5.3.4.3 (261) Direct sum of multiplets We have decomposed the rep that comes from a direct product of irreps into a direct sum of irreps. In literature, this direct sum of irreps is called the Clebsch-Gordan series (CG series). Since irreps can be labelled by their multiplets, we can use the notation for multiplets such as [j] to label the irreps. For two su(2) multiplets [j1 ] and [j2 ] with j1 ≥ j2 , the CG series read [j1 ] ⊗ [j2 ] = [j1 + j2 ] ⊕ [j1 + j2 − 1] ⊕ · · · ⊕ [j1 − j2 ] For example, we have for the fundamental multiplet of su(2) 1 1 ⊗ = [1] ⊕ [0], 2 2 1 1 1 3 1 ⊗ ⊗ = ⊕2 . 2 2 2 2 2 (262) (263) (264) A drawing technique for general reduction rules can be found in Pfeifer. 5.4 Cartan subalgebra The Cartan subalgebra is one-dimensional in σ3 . 5.4.1 Casimir operators There is only 1 Casimir operator, defined by C2 = σi σi ∝ J 2 , where Ji = 21 σi are the angular momentum operators. The su(2) subalgebra 49 5.4.2 Fundamental rep su(2) in 2D, su(3) in 3D, etc; for su(2), only 1 fundamental rep dual rep of su(2) has the same weight diagram, thus ρ ' ρ∗ for su(2). for su(3) not the case. antiquarks live in the dual space. p326, Greiner The su(3) Lie algebra and its reps 6 greiner chapter 7 pf p50-86 6.1 Generators and structure constants Gell-Mann’s matrices 6.2 6.2.1 Reps Fundamental and antifundamental reps fundamental rep (3) antifundamental rep 3̄ is the complex conjugate of the fundamental rep. 6.2.2 Tensor product rep 9-dimensional, reducible, can be decompose into 6 + 3̄ cartan-weyl basis weight = m, j = heighest weight weight lattice vs root lattice killing form simple roots Dynkin diagram representation = collection of weights Pi casimir operator 6.2.3 Multiplets The multiplet states are defined by |jmi states in the usual angular momentum language. in su3, 3 of 8 generators form an su2 subalgebra, other 5 form two doublets and a singlet. weyl group 6.2.4 Cartan subalgebra cartan subalbegra is the maximal abelian subalgebra. λ3 and λ8 generate a Cartan subalgebra. rank = 3. For any reps of su(3), the basis vector can be chosen to be the eigenvectors of (H1 , H2 ) = (T3 , T8 ). Ha |~ µi = µa |~ µi, a = 1, 2. The other 6 operators are ladder operators. Let us define E±~α1 ≡ T1 ± iT2 , (265) E±~α2 ≡ T4 ± iT5 , (266) E±~α3 ≡ T6 ± iT7 . (267) 50 Then [Hα , E±~α ] = α ~ α E±~α . (268) 2π α ~ 1 = (1, 0), α ~ 2 = cos π3 , sin π3 , α ~ 3 = cos 2π 3 , sin 3 . The ladder operators shift eigenvalues of Ha by the corresponding α ~ vector. rank is dim of Cartan subalgebra (2 for SU(3)) p166 figure 8.1 Kirillov The su(4) Lie algebra and its reps 7 chapter 11, Greiner su4, 60, 60, 90 degress!! pf p87-106 The generators of SU(4) can be written down from Eqs. (9)–(11). By convention, they are labeled in the following way 0 1 λ1 = 0 0 0 0 λ5 = i 0 0 0 λ9 = 0 1 0 0 λ13 = 0 0 1 0 0 0 0 0 0 0 0 0 0 0 −i 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 −i i 0 0 , λ2 = 0 0 0 0 0 0 0 0 0 0 0 0 , λ6 = 0 1 0 0 0 0 1 0 0 0 0 0 , λ10 = 0 0 0 0 i 0 0 0 0 0 0 0 , λ14 = 0 0 1 0 0 0 0 0 1 0 0 0 0 0 0 −1 0 0 0 0 0 0 , λ3 = 0 0 0 0 , λ4 = 1 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 −i 0 1 1 0 , λ7 = 0 i 0 0 , λ8 = √3 0 0 0 0 0 0 0 0 0 0 0 −i 0 0 0 0 0 0 0 0 0 1 0 0 0 0 , λ11 = 0 0 0 0 , λ12 = 0 0 0 0 0 0 0 1 0 0 0 i 1 0 0 0 0 0 1 0 1 0 0 0 0 . , λ15 = √ 0 −i 6 0 0 1 0 0 0 0 −3 i 0 1 0 0 0 0 0 , 0 0 0 0 1 0 0 −2 0 0 0 0 0 0 0 0 , 0 0 0 −i , 0 0 (269) There are 3 Casimir operators, of order 2, 3, and 451 . The quadratic Casimir operator C2 is given by C2 = λi λi + The nonzero structure constants fjkl and coefficients fjkl are given in figure (2). + Figure 2: The nonvanishing fjkl and fjkl for SU(4). su(4) has 3 su(2) subalgebras 8 Others 3. griener and muller fill it up, p456 for sl3 and sl4, go back to greiner after everything!!!!! 51 I will include higher orders of Casimir operators and the Cartan-Weyl basis later. 51 (270) example 4.22 p59 Kirillov, symmetric tensor, sym2 and gamma2, wiki su(3) torus, weight decomposition is now lattices. since not perfect equilateral triangle, want an angle. L Theorem 35: If ρ : su(3) → gl(V ) is a complex rep then V = Wk,l where Wk,l = {v ∈ V : ρ(D(θ1 , θ2 ))v} = ei(kθ1 +lθ2 v for k, l ∈ Z. sl(2, C) is a subalgebra of sl(3, C) from weight decomposition. sl(3, C) splits as a sum of sl(2, C) reps in 3 ways. Weyle symmetry, Weyl group isomorphic to S3 . Can use Weyl group and one highest weight to get the whole weight decomposition. Theorem 36: For every k, l ∈ N, there is an irrep Γk,l of su(3), unique up to isomorphism, whose weight diagram is the following. Moreover this is a complete list. classification -reflex using Weyle group -take the convex hull of these points to get the polygon p -take as weights the points in p of the form λ + r with r in the root lattice root lattice is a sublattice of the wight lattice, a linear combination of roots each dot is a weight space, being a subspace of the rep still need a way to get the dimension of the weight space (multiplicity), there is a algorithm for this. ⊗3 3 quarks = C3 Dual rep: ρ∗ : g → gl(V ∗ ) so that ρ∗ (a)v = v ∗ ρ(a)−1 . Dual space: 9 References Most useful references. 1. The Lie Algebra su(N) An Introduction, Pfeifer 2. Introduction to Lie Groups and Lie Algebras, Kirillov 3. Quantum Mechanics Symmetries, Greiner and Müller 4. Representation Theory and Quantum Mechanics, Miller 5. Quantum Theory, Groups, and Representations, Woit 6. Introduction to Lie Algebras and Representation Theory, Humpherys 7. Youtube videos, Jonathan Evans Others. 1. Introduction to Smooth Manifolds, Lee 2. Representation Theory, Fulton and Harris 3. Young Tableaux, Fulton 4. Group Theory in Physics, Tung 5. Group Theory: Birdtracks, Lie’s, and Exceptional Groups, Cvitanović 6. Lecture notes on Hadron Physics, Eichmann 7. Shankar 52