9.2 Lie groups, Lie algebras and representations
As it becomes evident from our informal discussion in the previous section, the concepts of Lie groups, Lie algebras and their representations are crucial to understand the structure of the standard model of particle physics. The aim of this section is to give a very brief (and hence necessarily incomplete) introduction to these rich topics, focusing mostly on the specific examples that we require to set up the standard model.
Informally speaking, Lie groups are used to describe continuous families of transformations that depend smoothly on their parameters. The general mathematical definition combines concepts from differential geometry and group theory, and it reads as follows:
-
Definition 9.1 (Lie group). A Lie group is a smooth manifold \(G\) that is equipped with a smooth map \(G\times G\to G\,,~(g^\prime ,g)\mapsto g^\prime \,g\) (called group multiplication), an element \(e\in G\) (called identity element) and a smooth map \(G\to G\,,~g\mapsto g^{-1}\) (called inversion). These structures have to satisfy the axioms of a group, i.e.
-
(i) Associativity: \((g^{\prime \prime }\,g^\prime )\,g = g^{\prime \prime }\,(g^\prime \,g)\) for all \(g,g^\prime ,g^{\prime \prime }\in G\)
-
(ii) Identity property: \(g\,e = g = e\,g\) for all \(g\in G\)
-
(iii) Inverse property: \(g^{-1}\,g = e = g\,g^{-1}\) for all \(g\in G\)
A Lie group \(G\) is called Abelian if the group multiplication is commutative, i.e. \(g^\prime \,g = g\,g^\prime \) for all \(g,g^\prime \in G\), and otherwise it is called non-Abelian.
-
This definition is very general and hence it captures many relevant examples. For instance, one can endow the (proper and orthochronous) Lorentz group \(\mathsf {SO}_0(d-1,1)\), the spin group \(\mathsf {Spin}(d-1,1)\) and the Poincaré group with a canonical manifold structure such that they become Lie groups. In the context of the standard model of particle physics, we do not necessarily need this general definition of a Lie group and it suffices to understand the following families of examples.
-
Example 9.2 (The Lie groups \(\U (n)\) and \(\SU (n)\)). Let \(n\geq 1\) be a positive integer. The unitary group of degree \(n\) is defined as the submanifold
\(\seteqnumber{0}{9.}{0}\)\begin{equation} \mathsf {U}(n)\,:=\,\Big \{U\in \mathrm {Mat}_{n\times n}(\bbC )\,:\, U^\dagger \,U =\mathbf {1} = U\,U^\dagger \Big \}\,\subseteq \, \mathrm {Mat}_{n\times n}(\bbC ) \end{equation}
that consists of all unitary \(n\times n\)-matrices, together with the group multiplication \(U^\prime \,U\) given by matrix multiplication, the identity element \(e=\mathbf {1}\) given by the identity matrix and the inversion \(U^{-1}\) given by the inverse matrix. (Note that the latter is equal to the adjoint matrix \(U^{-1} = U^\dagger \) because we consider unitary matrices.) Setting \(n=1\), we observe that the circle group \(\mathsf {U}(1) = \{U\in \bbC \,:\,U^\ast \,U = 1\}\subseteq \bbC \) is the unitary group of degree \(1\). The special unitary group of degree \(n\) is defined as the Lie subgroup
\(\seteqnumber{0}{9.}{1}\)\begin{equation} \mathsf {SU}(n)\,:=\,\Big \{U\in \mathrm {Mat}_{n\times n}(\bbC )\,:\, U^\dagger \,U =\mathbf {1} = U\,U^\dagger ~~ \text {and}~~ \det U =1\Big \}\,\subseteq \, \mathsf {U}(n) \end{equation}
that consists of all unitary \(n\times n\)-matrices with determinant \(1\). Since the multiplication of matrices of size \(n\geq 2\) is non-commutative, we observe that \(\mathsf {U}(n)\) and \(\mathsf {SU}(n)\) are non-Abelian Lie groups for \(n\geq 2\). In contrast to this, circle group \(\mathsf {U}(1)\) is an Abelian Lie group.
Since a Lie group \(G\) describes by design a notion of smooth families of transformations, one should expect that there is an associated notion of infinitesimal transformations that can be obtained via a suitable differentiation procedure. This is indeed the case and can be formalized using the concept of tangent spaces from differential geometry. More precisely, the infinitesimal transformations are characterized by the tangent space \(\g :=T_eG\) at the identity element \(e\in G\), which can be endowed with a Lie algebra structure \([\,\cdot \,,\,\cdot \,]:\g \times \g \to \g \) that is induced by the Lie group structure on \(G\). We have already seen some examples of Lie algebras in this module, but we have never written down a general definition. So let us do this now.
-
Definition 9.3 (Lie algebra). A (real) Lie algebra is a real vector space \(\g \) together with a bilinear map \([\,\cdot \,,\,\cdot \,] : \g \times \g \to \g \) (called Lie bracket) that satisfies the following axioms:
-
(i) Antisymmetry: \([X,Y] = - [Y,X]\) for all \(X,Y\in \g \)
-
(ii) Jacobi identity: \([X,[Y,Z]] + [Y,[Z,X]]+[Z,[X,Y]] =0\) for all \(X,Y,Z\in \g \)
A Lie algebra \(\g \) is called Abelian if the Lie bracket is trivial, i.e. \([X,Y]=0\) for all \(X,Y\in \g \), and otherwise it is called non-Abelian.
-
Let us illustrate this concept by working out the Lie algebras associated with the unitary and special unitary groups.
-
Example 9.4 (The Lie algebras \(\mathfrak {u}(n)\) and \(\mathfrak {su}(n)\)). Consider the unitary Lie group \(\U (n)\) from Example 9.2, whose elements are unitary \(n\times n\)-matrices \(U\). In the vicinity of the identity element \(\mathbf {1}\in \U (n)\), we can write such matrices in exponential form \(U = e^{X}\), where \(X\) is some small \(n\times n\)-matrix. A first-order Taylor expansion in \(X\) of the unitarity condition \(U^\dagger \,U = \mathbf {1}\) yields
\(\seteqnumber{0}{9.}{2}\)\begin{flalign} U^\dagger \,U = e^{X^\dagger }\,e^X = \mathbf {1} + X^\dagger + X +\mathcal {O}(X^2) \stackrel {!}{=}\mathbf {1}\quad , \end{flalign} which means that the exponent must be an anti-Hermitian matrix \(X^\dagger = -X\). (Note that the second condition \(U\,U^\dagger = \mathbf {1}\) yields the same anti-Hermiticity condition for \(X\).) This implies that the Lie algebra of \(\U (n)\) is given by the vector space
\(\seteqnumber{1}{9.4}{0}\)\begin{equation} \mathfrak {u}(n)\,:=\, \Big \{X\in \mathrm {Mat}_{n\times n}(\bbC )\,:\, X^\dagger = -X\Big \} \end{equation}
of anti-Hermitian \(n\times n\)-matrices, on which the canonically defined Lie bracket is given by the matrix commutator
\(\seteqnumber{1}{9.4}{1}\)\begin{equation} [\,\cdot \,,\,\cdot \,] \,:\, \mathfrak {u}(n)\times \mathfrak {u}(n)\longrightarrow \mathfrak {u}(n)~,~~(X,Y)\longmapsto [X,Y] \,:=\,X\,Y - Y\,X \quad . \end{equation}
Note that \([X,Y]\in \mathfrak {u}(n)\) is indeed anti-Hermitian for all \(X,Y\in \mathfrak {u}(n)\), which is shown by a short calculation
\(\seteqnumber{0}{9.}{4}\)\begin{flalign} [X,Y]^\dagger = [Y^\dagger ,X^\dagger ] = [Y,X] = -[X,Y]\quad . \end{flalign} For the special unitary Lie group \(\SU (n)\subseteq \U (n)\) there is the extra condition that \(\det U=1\) must be \(1\). Using the relationship \(\det (e^X) = e^{\mathrm {Tr}(X)}\) between determinant and trace, we find that the Lie algebra of \(\SU (n)\) is given by the vector space
\(\seteqnumber{1}{9.6}{0}\)\begin{equation} \mathfrak {su}(n)\,:=\, \Big \{X\in \mathrm {Mat}_{n\times n}(\bbC )\,:\, X^\dagger = -X~~\text {and}~~\mathrm {Tr}(X)=0\Big \} \end{equation}
of anti-Hermitian and trace-free \(n\times n\)-matrices, on which the canonically defined Lie bracket is given again by the matrix commutator
\(\seteqnumber{1}{9.6}{1}\)\begin{equation} [\,\cdot \,,\,\cdot \,] \,:\, \mathfrak {su}(n)\times \mathfrak {su}(n)\longrightarrow \mathfrak {su}(n)~,~~(X,Y)\longmapsto [X,Y] \,:=\,X\,Y - Y\,X \quad . \end{equation}
Note that \([X,Y]\in \mathfrak {su}(n)\) is indeed trace-free for all \(X,Y\in \mathfrak {su}(n)\), which is shown by using the cyclicity property of the trace
\(\seteqnumber{0}{9.}{6}\)\begin{flalign} \mathrm {Tr}\big ([X,Y]\big ) = \mathrm {Tr}\big (X\,Y\big ) -\mathrm {Tr}\big (Y\,X\big ) = \mathrm {Tr}\big (X\,Y\big ) -\mathrm {Tr}\big (X\,Y\big ) =0\quad . \end{flalign} We observe that \(\mathfrak {u}(n)\) and \(\mathfrak {su}(n)\) are non-Abelian Lie algebras for \(n\geq 2\), and that the Lie algebra \(\mathrm {u}(1) = \ii \,\bbR \) of the circle group is Abelian and \(1\)-dimensional.
-
Remark 9.5 (Structure constants of a Lie algebra). Given any Lie algebra \(\g \), we can pick a basis \(\{X_a\in \g \}\) of its underlying vector space. Since the Lie bracket is bilinear, it is completely characterized by its values \([X_a,X_b]\in \g \) on the basis elements, which can be expanded in the chosen basis as
\(\seteqnumber{0}{9.}{7}\)\begin{flalign} [X_a,X_b] = f_{ab}^c\,X_c\quad . \end{flalign} The expansion coefficients \(f_{ab}^c\in \bbR \) are called the structure constants of the Lie algebra in the chosen basis. At the level of the structure constants, antisymmetry of the Lie bracket reads as \(f_{ab}^c = -f_{ba}^c\) and the Jacobi identity translates to
\(\seteqnumber{0}{9.}{8}\)\begin{flalign} f_{ad}^e\,f_{bc}^d +f_{bd}^e\,f_{ca}^d+f_{cd}^e\,f_{ab}^d =0\quad . \end{flalign} This description of Lie algebras in terms of a basis is what you’ll typically see in physics textbooks. Since using a basis \(X_a\) introduces further indices and thereby complicates notation, we will present most of our constructions in this chapter in a basis independent way.
To get a better feeling for how picking a basis works in practice, let us consider the Lie algebra \(\mathfrak {su}(2)\) of anti-Hermitian and trace-free \(2\times 2\)-matrices. A basis for \(\mathfrak {su}(2)\) is given by the following matrices
\(\seteqnumber{0}{9.}{9}\)\begin{flalign} X_1 = -\frac {\ii }{2}\begin{pmatrix} 0 & 1\\ 1 & 0 \end {pmatrix}~~,\quad X_2 = -\frac {\ii }{2} \begin{pmatrix} 0 & -\ii \\ \ii & 0 \end {pmatrix}~~,\quad X_3 =-\frac {\ii }{2} \begin{pmatrix} 1 & 0\\ 0 & -1 \end {pmatrix}\quad , \end{flalign} which you probably recognize as the Pauli matrices rescaled by a factor of \(-\frac {\ii }{2}\). Computing the commutators, one finds
\(\seteqnumber{0}{9.}{10}\)\begin{flalign} [X_a,X_b] = \epsilon _{abc}\,X_c\quad , \end{flalign} i.e. the structure constants \(f_{ab}^c = \epsilon _{abc}\) are given by the epsilon-tensor. The reason why there is no factor \(\ii \) on the right-hand side, as you would expect by comparing this with the \(\mathfrak {su}(2)\) Lie algebra \([L_a,L_b] = \ii \,\epsilon _{abc}\,L_c\) from your quantum mechanics module, is that we have taken different conventions for exponentials. Our convention was to write \(U = e^X\), leading to anti-Hermitian exponents \(X\), while in quantum mechanics people typically work with \(U= e^{-\ii \,L}\), leading to Hermitian exponents \(L\). Note that one can pass from one convention to the other by setting \(L_a = \ii \,X_a\), which takes care of the factor \(\ii \) in the Lie bracket relations.
By design, a Lie group \(G\) describes an abstract notion of transformations and its Lie algebra \(\g \) describes an abstract notion of infinitesimal transformations. To let these transformations act concretely on other objects, such as a vector space \(V\), one introduces the concept of representations. To understand the following definition, we observe that associated with every vector space \(V\) is the Lie group \(\mathsf {GL}(V) \subseteq \mathrm {Lin}(V,V)\) of invertible linear maps \(L : V\to V\) from \(V\) to itself, with group multiplication \(L^\prime \circ L\) given by composition of linear maps, identity element \(\mathrm {id}\) given by the identity map and inversion \(L^{-1}\) given by the inverse linear map.
-
Definition 9.6 (Representation of a Lie group). A (linear) representation of a Lie group \(G\) on a real or complex vector space \(V\) is a smooth group homomorphism
\(\seteqnumber{0}{9.}{11}\)\begin{flalign} \rho \,:\, G \longrightarrow \mathsf {GL}(V)~,~~g\longmapsto \rho (g)\quad , \end{flalign} i.e. it preserves the group multiplication \(\rho (g^\prime \,g) = \rho (g^\prime )\circ \rho (g)\), for all \(g,g^\prime \in G\). (From this property one can deduce that the identity element \(\rho (e) = \mathrm {id}\) and inverses \(\rho (g^{-1}) = \big (\rho (g)\big )^{-1}\), for all \(g\in G\), are preserved too.) In the case where \(V\) is a complex vector space endowed with a sesquilinear inner product \(\langle \,\cdot \,,\,\cdot \,\rangle : V\times V\to \bbC \), one says that a representation \(\rho \) is unitary if it preserves the inner product in the sense that
\(\seteqnumber{0}{9.}{12}\)\begin{flalign} \big \langle \rho (g)(v), \rho (g)(v^\prime )\big \rangle \,=\,\langle v,v^\prime \rangle \quad , \end{flalign} for all \(g\in G\) and \(v,v^\prime \in V\). The notation \(\rho (g)(v)\in V\) means the application of the linear map \(\rho (g) : V\to V\) on the element \(v\in V\).
The classification of linear representations of Lie groups is an interesting and non-trivial problem that is studied in the mathematical discipline called representation theory. Luckily, for understanding the standard model of particle physics, it suffices to know only some very specific representations of the special unitary group \(\mathsf {SU}(n)\) and the representations of the circle group \(\mathsf {U}(1)\).
-
Example 9.7 (Important representations of \(\SU (n)\)). We list some basic representations of \(\SU (n)\) that are needed for constructing the standard model of particle physics.
-
1. Trivial representation: The trivial representation on the \(1\)-dimensional complex vector space \(\bbC \) is given by the smooth group homomorphism
\(\seteqnumber{0}{9.}{13}\)\begin{equation} \rho _{\mathrm {triv}} \,:\,\SU (n)\longrightarrow \mathsf {GL}(\bbC )~,~~U\longmapsto \rho _{\mathrm {triv}}(U) = \mathrm {id} \end{equation}
that sends every \(U\in \SU (n)\) to the identity map. This representation is unitary with respect to the standard inner product \(\langle v,v^\prime \rangle = v^\ast \,v^\prime \) on \(\bbC \).
-
2. Defining representation: Consider the standard \(n\)-dimensional complex vector space \(\bbC ^n\). Recall from linear algebra that \(\mathrm {Lin}(\bbC ^n,\bbC ^n)\cong \mathrm {Mat}_{n\times n}(\bbC )\) can be identified with the \(n\times n\)-matrices with complex entries, hence \(\mathsf {GL}(\bbC ^n)\) can be identified with the Lie group of invertible \(n\times n\)-matrices. Explicitly, each \(n\times n\)-matrix \(A\in \mathrm {Mat}_{n\times n}(\bbC )\) defines via matrix multiplication a linear map \(\bbC ^n\to \bbC ^n\,,~v\mapsto A\,v\), and every linear map can be written uniquely in this way. The defining representation
\(\seteqnumber{0}{9.}{14}\)\begin{equation} \rho _{\mathrm {def}}\,:\,\mathsf {SU}(n) \longrightarrow \mathsf {GL}(\bbC ^n)~,~~U\longmapsto \rho _{\mathrm {def}}(U) = U \end{equation}
is then simply given by the embedding of the unitary matrices with determinant \(1\) into all invertible matrices. Endowing \(\bbC ^n\) with its standard inner product \(\langle v,v^\prime \rangle := v^\dagger \,v^\prime \), one easily checks that this representation is unitary
\(\seteqnumber{0}{9.}{15}\)\begin{flalign} \big \langle \rho _{\mathrm {def}}(U)(v) , \rho _{\mathrm {def}}(U)(v^\prime )\big \rangle =\big \langle U\,v , U\, v^\prime \big \rangle = v^\dagger \,U^\dagger \,U\,v^\prime = v^\dagger \,v^\prime = \langle v,v^\prime \rangle \quad . \end{flalign}
-
3. Adjoint representation: Every Lie group \(G\) has a canonical linear representation on its Lie algebra \(\g :=T_e G\), which is called the adjoint representation. For \(\SU (n)\), the adjoint representation
\(\seteqnumber{1}{9.17}{0}\)\begin{equation} \rho _{\mathrm {ad}} \,:\,\mathsf {SU}(n)\longrightarrow \mathsf {GL}(\mathfrak {su}(n)) \end{equation}
is given element-wise in terms of matrix multiplication by
\(\seteqnumber{1}{9.17}{1}\)\begin{equation} \rho _{\mathrm {ad}}(U)(X)\,:=\, U\,X\,U^{-1} \,=\, U\,X\,U^{\dagger }\quad , \end{equation}
where in the second step we used that \(U\) is unitary, i.e. \(U^{-1} = U^\dagger \). This is indeed well-defined: One easily checks that the matrix \(\rho _{\mathrm {ad}}(U)(X)\) is anti-Hermitian
\(\seteqnumber{0}{9.}{17}\)\begin{flalign} \big (\rho _{\mathrm {ad}}(U)(X)\big )^\dagger = \big (U\,X\,U^{\dagger }\big )^\dagger = U\,X^\dagger \,U^\dagger = -\rho _{\mathrm {ad}}(U)(X) \end{flalign} and trace-free
\(\seteqnumber{0}{9.}{18}\)\begin{flalign} \mathrm {Tr}\big (\rho _{\mathrm {ad}}(U)(X)\big ) = \mathrm {Tr}\big (U\,X\,U^\dagger \big ) = \mathrm {Tr}\big (U^\dagger \,U\,X\big ) = \mathrm {Tr}\big (X\big ) =0\quad , \end{flalign} hence \(\rho _{\mathrm {ad}}(U)(X)\in \mathfrak {su}(n)\) defines an element of the Lie algebra. One also checks easily that \(\rho _{\mathrm {ad}}\) defines a representation, which follows from the short calculation
\(\seteqnumber{0}{9.}{19}\)\begin{flalign} \rho _{\mathrm {ad}}(U^\prime \,U)(X) = (U^\prime \,U) \,X\,(U^\prime \,U)^\dagger =U^\prime \, \big (U \,X\,U^\dagger \big )\, U^{\prime \,\dagger }= \rho _{\mathrm {ad}}(U^\prime )\big (\rho _{\mathrm {ad}}(U)(X)\big )\quad . \end{flalign}
Note that the representations above also make sense when we replace \(\SU (n)\) by the unitary Lie group \(\U (n)\).
-
-
Example 9.8 (Representations of \(\mathsf {U}(1)\)). It is quite easy to write down explicit examples of unitary linear representations of the circle group \(\mathsf {U}(1) = \{U\in \bbC \,:\,U^\ast \,U = 1\}\subseteq \bbC \) on the \(1\)-dimensional complex vector space \(\bbC \). In this case we have that \(\mathsf {GL}(\bbC ) = \{U\in \bbC \,:\,U\neq 0\} = \bbC \setminus \{0\}\) is the complex plane minus the origin. For each integer \(k\in \bbZ \), we can define a unitary representation
\(\seteqnumber{0}{9.}{20}\)\begin{equation} \rho ^{(k)}\,:\,\mathsf {U}(1)\longrightarrow \mathsf {GL}(\bbC )~,~~U\longmapsto U^k\quad \end{equation}
by taking \(U\) to the power \(k\). Indeed, one easily checks that \(\rho ^{(k)}(U^\prime \,U) = (U^\prime \,U)^k = {U^\prime }^k\,U^k = \rho ^{(k)}(U^\prime )\,\rho ^{(k)}(U)\), for all \(U,U^\prime \in \mathsf {U}(1)\). With some more efforts, which won’t be discuss here, one can show that these are all irreducible linear representations of \(\mathsf {U}(1)\), i.e. every other linear representation can be build from these \(\rho ^{(k)}\)’s.
-
Remark 9.9 (New representations from old ones). There exist general techniques to construct new representations out of given ones. A detailed discussion of these methods goes beyond the scope of this module, but as a taster I would like to sketch some of the basic ideas.
-
(i) Tensor product representations: Let \(G\) be a Lie group and suppose that we have given two vector spaces \(V\) and \(W\) with representations \(\rho _V\) and \(\rho _W\), respectively. Then the tensor product representation is defined by
\(\seteqnumber{0}{9.}{21}\)\begin{flalign} \rho _{V\otimes W} \,:\,G\longrightarrow \mathsf {GL}(V\otimes W) ~,~~g\longmapsto \rho _V(g)\otimes \rho _W(g)\quad . \end{flalign} You probably have seen simple examples of such tensor product representations when you’ve studied spin in quantum mechanics, which are unitary \(\mathsf {SU}(2)\) representations. More concretely, a non-relativistic spin \(\frac {1}{2}\) particle is described by the defining representation of \(\mathsf {SU}(2)\), which in physics is typically written in terms of the states \(\ket {\,{\uparrow }\,},\ket {\,{\downarrow }\,}\in \bbC ^2\). The tensor product of this representation with itself is described by the tensor product states \(\ket {\,{\uparrow \uparrow }\,},\ket {\,{\uparrow \downarrow }\,},\ket {\,{\downarrow \uparrow }\,}, \ket {\,{\downarrow \downarrow }\,} \in \bbC ^2\otimes \bbC ^2\). A standard exercise in quantum mechanics is then to show that this tensor product representation can be decomposed into a spin \(1\) and a spin \(0\) representation by using Clebsch-Gordan coefficients.
-
(ii) Dual representations: Let \(G\) be a Lie group and suppose that we have given a representation \(\rho \) on a vector space \(V\). Recall that the dual vector space \(V^\ast :=\mathrm {Lin}(V,\mathbb {K})\) is given by the vector space of linear maps \(f:V\to \mathbb {K}\) from \(V\) to the \(1\)-dimensional vector space \(\mathbb {K}\), i.e. \(\mathbb {K} = \bbR \) in the real case and \(\mathbb {K} = \bbC \) in the complex case. Then the dual representation
\(\seteqnumber{1}{9.23}{0}\)\begin{flalign} \rho ^\ast \,:\,G\longrightarrow \mathsf {GL}(V^\ast ) \end{flalign} is defined element-wise by
\(\seteqnumber{1}{9.23}{1}\)\begin{flalign} \rho ^\ast (g)(f)\,:=\, f\circ \rho (g^{-1})\quad , \end{flalign} for all \(g\in G\) and \(f\in V^\ast \).
-
There is also a concept of representations for Lie algebras on vector spaces. To understand the following definition, we observe that associated with every vector space \(V\) is the Lie algebra \(\mathfrak {gl}(V) := \mathrm {Lin}(V,V)\) of linear maps \(L : V\to V\) from \(V\) to itself, with Lie bracket \([L,L^\prime ]:= L\circ L^\prime - L^\prime \circ L\) given by the commutator.
-
Definition 9.10 (Representation of a Lie algebra). A representation of a Lie algebra \(\mathfrak {g}\) on a real or complex vector space \(V\) is a Lie algebra homomorphism
\(\seteqnumber{0}{9.}{23}\)\begin{flalign} \underline {\rho }\,:\, \mathfrak {g} \longrightarrow \mathfrak {gl}(V)~,~~X\longmapsto \underline {\rho }(X)\quad , \end{flalign} i.e. a linear map that preserves the Lie bracket
\(\seteqnumber{0}{9.}{24}\)\begin{flalign} \underline {\rho }\big ([X,Y]\big ) = \big [\underline {\rho }(X),\underline {\rho }(Y)\big ]\quad , \end{flalign} for all \(X,Y\in \mathfrak {g}\). In the case where \(V\) is a complex vector space endowed with a sesquilinear inner product \(\langle \,\cdot \,,\,\cdot \,\rangle : V\times V\to \bbC \), one says that a representation \(\underline {\rho }\) is unitary if it preserves the inner product in the sense that
\(\seteqnumber{0}{9.}{25}\)\begin{flalign} \big \langle \underline {\rho }(X)(v), v^\prime \big \rangle + \big \langle v, \underline {\rho }(X)(v^\prime )\big \rangle \,=\,0\quad , \end{flalign} for all \(X\in \mathfrak {g}\) and \(v,v^\prime \in V\). The notation \(\underline {\rho }(X)(v)\in V\) means the application of the linear map \(\underline {\rho }(X) : V\to V\) on the element \(v\in V\).
One can show that every linear representation \(\rho \) of a Lie group \(G\) on a vector space \(V\) induces a representation \(\underline {\rho }\) of its underlying Lie algebra \(\g =T_eG\) on the same \(V\). Instead of explaining this construction in full generality, we give some relevant examples.
-
Example 9.11 (Important representations of \(\mathfrak {su}(n)\)). Before going into the examples, let me sketch the basic idea how one can construct Lie algebra representations from linear Lie group representations: Given any linear representation \(\rho : \SU (n)\to \mathsf {GL}(V)\) of the Lie group \(\SU (n)\), we can consider the linear map \(\rho (e^{X}) : V\to V\) that is associated to \(U=e^X\in \SU (n)\) for a Lie algebra element \(X\in \mathfrak {su}(n)\). Using a first-order Taylor expansion in \(X\), we define
\(\seteqnumber{0}{9.}{26}\)\begin{flalign} \rho (e^{X}) =: \mathrm {id} + \underline {\rho }(X) + \mathcal {O}(X^2)\quad . \end{flalign} From the Baker-Campbell-Hausdorff formula \(e^{X}\,e^Y = e^{X+Y + \frac {1}{2}[X,Y] + \cdots }\) and the property that \(\rho \) is a Lie group representation, one shows that \(\underline {\rho }\) is a Lie algebra representation, i.e.
\(\seteqnumber{0}{9.}{27}\)\begin{flalign} \underline {\rho }\big ([X,Y]\big )\,= \, \big [\underline {\rho }(X),\underline {\rho }(Y)\big ]\quad , \end{flalign} for all \(X,Y\in \mathfrak {su}(n)\). The Lie algebra representations associated with the \(\SU (n)\) representations listed in Example 9.7 read concretely as follows:
-
1. Trivial representation: By definition, we have that
\(\seteqnumber{0}{9.}{28}\)\begin{flalign} \rho _{\mathrm {triv}}(e^X) = \mathrm {id}\quad , \end{flalign} hence \(\underline {\rho _{\mathrm {triv}}}(X) = 0\) for all \(X\in \mathfrak {su}(n)\). This Lie algebra representation is unitary.
-
2. Defining representation: Let us compute
\(\seteqnumber{0}{9.}{29}\)\begin{flalign} \rho _{\mathrm {def}}(e^X) = e^X = \mathrm {id} + X +\mathcal {O}(X^2)\quad , \end{flalign} hence we have that \(\underline {\rho _{\mathrm {def}}}(X) = X\) for all \(X\in \mathfrak {su}(n)\). This Lie algebra representation is unitary when \(\bbC ^n\) is endowed with its standard inner product \(\langle v,v^\prime \rangle := v^\dagger \,v^\prime \). Indeed,
\(\seteqnumber{0}{9.}{30}\)\begin{flalign} \big \langle \underline {\rho _{\mathrm {def}}}(X)(v) , v^\prime \big \rangle + \big \langle v , \underline {\rho _{\mathrm {def}}}(X)(v^\prime )\big \rangle = v^\dagger \, X^\dagger \,v^\prime + v^\dagger \, X\,v^\prime =0\quad , \end{flalign} for all \(X\in \mathfrak {su}(n)\) and \(v,v^\prime \in \bbC ^n\), where in the last step we used that \(X^\dagger = -X\) is anti-Hermitian.
-
3. Adjoint representation: Let us compute
\(\seteqnumber{0}{9.}{31}\)\begin{flalign} \rho _{\mathrm {ad}}(e^X)(Y)\,:=\, e^{X}\,Y\,e^{-X} \,=\,Y + X\,Y - Y\,X +\mathcal {O}(X^2) \,=\, Y + [X,Y] +\mathcal {O}(X^2)\quad , \end{flalign} hence we have that \(\underline {\rho _{\mathrm {ad}}}(X)(Y) = [X,Y]\) is given by the Lie bracket for all \(X,Y\in \mathfrak {su}(n)\).
Note that the representations above also make sense when we replace \(\SU (n)\) by the unitary Lie group \(\U (n)\).
-