Fact-checked by Grok 2 weeks ago

Simple Lie group

A simple Lie group is a connected, non-abelian that possesses no nontrivial connected Lie subgroups. These groups form a fundamental class of s, which are manifolds equipped with a group structure where the group operations of multiplication and inversion are maps. The defining property of simplicity ensures that the associated —the at the endowed with the Lie bracket derived from the group's —is itself simple, meaning it has no nontrivial ideals. The classification of simple Lie groups, achieved through the study of their Lie algebras, reveals a finite number of exceptional cases alongside infinite families of classical types. Over the complex numbers, the simple Lie algebras (and thus the corresponding simply connected simple Lie groups) are categorized by their root systems, corresponding to Dynkin diagrams labeled as A_n (n ≥ 1) for sl(n+1, ℂ), B_n (n ≥ 2) for so(2n+1, ℂ), C_n (n ≥ 3) for sp(n, ℂ), D_n (n ≥ 4) for so(2n, ℂ), and the exceptional types G_2, F_4, E_6, E_7, E_8. This classification, originally due to and in the late 19th and early 20th centuries, provides a complete enumeration and underpins the structural theory of these groups. For real simple Lie groups, the picture is richer, involving various real forms of the complex ones, but the compact real forms align directly with the complex classification. Simple Lie groups play a pivotal role in modern and physics, serving as building blocks for semisimple Lie groups via direct products and influencing areas such as , where irreducible representations decompose under actions. In physics, they model symmetries in quantum field theories; for instance, SU(3) (from A_2) underlies the strong nuclear force in , while SO(3) (from B_1) describes rotational symmetries in classical and . Their study also extends to , through connections to symmetric spaces and homogeneous manifolds, highlighting their interdisciplinary significance.

Definition and Basics

Formal Definition

A Lie group G is simple if it is connected and non-abelian, and possesses no nontrivial connected Lie subgroups other than the trivial subgroup \{e\} and G itself. This definition captures the indivisibility of the group's structure under continuous subgroups, excluding one-dimensional abelian cases which are not considered . Simple Lie groups are defined to be connected by convention, as the classification of disconnected Lie groups involves additional components handled separately. The term "simple" in this context pertains specifically to the Lie group structure and its continuous symmetries, distinct from the notion of simple groups in finite group theory, where no normal subgroups exist whatsoever. A fundamental result from the structure theory of Lie groups, developed by Killing, Cartan, and others, states that a connected Lie group is simple if and only if its Lie algebra is simple. For real Lie groups, this means a simple real Lie algebra, whose complexification is a simple complex Lie algebra. This allows the classification of simple Lie groups to proceed via the classification of simple complex Lie algebras. This correspondence underpins the classification of simple Lie groups through their associated algebras.

Equivalent Formulations

Simple Lie groups admit several equivalent characterizations that highlight their structural rigidity and connection to their underlying s. One key formulation identifies a simple Lie group G as isomorphic to the group \operatorname{Ad}(G) = G / Z(G) of a \mathfrak{g}, where Z(G) denotes of G. In this form, the group has trivial , meaning Z(G) = \{e\}, and the Lie algebra is realized as \mathfrak{g} / z(\mathfrak{g}) with z(\mathfrak{g}) = \{0\}, emphasizing the absence of central elements beyond the . This perspective arises from the , where the group acts by conjugation on itself, and is fundamental for understanding inner automorphisms in semisimple structures. Another equivalent view focuses on the center and covering properties: simple Lie groups possess discrete centers, which are finite for compact cases, ensuring that the only continuous central elements are trivial. The universal covering group \tilde{G} of a connected simple Lie group G inherits the simple Lie algebra, but the quotient by the discrete center yields the centerless adjoint form, providing a canonical centerless realization of the structure. This formulation underscores the role of the center in distinguishing covering groups while preserving simplicity at the infinitesimal level. Simple Lie groups also evade nontrivial Levi decompositions, possessing no nontrivial solvable subgroups. Any connected subgroup N of G corresponds to an ideal in the \mathfrak{g}, but since \mathfrak{g} is simple, the only such ideals are \{0\} and \mathfrak{g} itself, implying no proper connected subgroups exist. subgroups, if present, must lie in and thus be solvable, but the simplicity condition restricts them to triviality in the adjoint case; overall, this ensures the of G—the maximal solvable subgroup—is trivial. A precise algebraic characterization is that for a simple Lie group G, the derived subgroup G' = [G, G]—generated by all commutators [g, h] = g h g^{-1} h^{-1} for g, h \in G—coincides with G itself. This equality reflects the perfect nature of G, where the abelianization G / G' is trivial, mirroring the property of the Lie algebra where [\mathfrak{g}, \mathfrak{g}] = \mathfrak{g}. Such formulations collectively affirm the indecomposability and non-abelian perfection inherent to simple Lie groups.

Notation Conventions

In the study of simple Lie groups, standard notation for their associated complex simple Lie algebras follows the Killing-Cartan classification, where the classical series are denoted as follows: the type A_n corresponds to the \mathrm{SL}(n+1, \mathbb{C}) or its \mathfrak{sl}(n+1, \mathbb{C}) for n \geq 1; type B_n to the odd \mathrm{SO}(2n+1, \mathbb{C}) or \mathfrak{so}(2n+1, \mathbb{C}) for n \geq 1; type C_n to the \mathrm{Sp}(n, \mathbb{C}) or \mathfrak{sp}(n, \mathbb{C}) for n \geq 2, where \mathrm{Sp}(n, \mathbb{C}) acts on \mathbb{C}^{2n}; and type D_n to the even \mathrm{SO}(2n, \mathbb{C}) or \mathfrak{so}(2n, \mathbb{C}) for n \geq 3. The exceptional types are E_6, E_7, E_8, F_4, and G_2, each denoting a unique simply connected simple Lie group (up to ) with the corresponding . For real forms of these groups, notation distinguishes non-compact real structures from the case; for instance, in the A-series, the indefinite special \mathrm{SU}(p, q) (with p + q = n+1) provides a non-compact real form of the group \mathrm{SL}(n+1, \mathbb{C}), preserving a Hermitian form of (p, q). Similar conventions apply to other series, such as \mathrm{SO}(p, q) for B- and D-types, emphasizing the of the . Dynkin labels arise in the context of root systems for these Lie algebras, where the simple roots are denoted \alpha_1, \alpha_2, \dots, \alpha_l, forming a basis for the lattice; these labels a_i = \langle \lambda, \alpha_i^\vee \rangle (with \alpha_i^\vee the coroot) specify dominant weights \lambda in irreducible representations. Throughout this article, the of a simple Lie algebra (or its corresponding group) is denoted by l, equal to the dimension of a Cartan subalgebra or the number of simple roots.

Algebraic Foundations

Simple Lie Algebras

A \mathfrak{g} over a of characteristic zero is defined to be if it is non-abelian and admits no nontrivial proper ideals. This means that the only ideals of \mathfrak{g} are \{0\} and \mathfrak{g} itself, ensuring that \mathfrak{g} cannot be decomposed into smaller Lie subalgebras in a nontrivial way under the Lie bracket operation. Simple Lie algebras form the fundamental indecomposable building blocks in the structure theory of Lie algebras. A key property of simple Lie algebras is their role in the decomposition of more general structures: a Lie algebra is semisimple if and only if it is a direct sum of simple Lie algebras. In this context, simplicity implies the absence of ideals, which distinguishes simple algebras from semisimple ones that may consist of multiple simple components. This direct sum decomposition highlights how semisimple Lie algebras, which have no nonzero solvable ideals, reduce to the study of their simple summands. The Killing form provides an important invariant for simple Lie algebras. Defined as B(X, Y) = \operatorname{tr}(\operatorname{ad}_X \operatorname{ad}_Y) for X, Y \in \mathfrak{g}, where \operatorname{ad} denotes the , the Killing form is a that is non-degenerate on simple Lie algebras. This non-degeneracy follows from the semisimplicity of simple algebras and serves as a criterion to distinguish them from solvable or nilpotent ones, where the form may be degenerate. Fundamental theorems establish the deep connection between simple Lie algebras and Lie groups. Lie's third theorem asserts that every finite-dimensional Lie algebra over \mathbb{R} or \mathbb{C} is the Lie algebra of some , providing a bridge from the algebraic to the analytic setting. Complementing this, Ado's theorem guarantees that every finite-dimensional over a of characteristic zero admits a faithful finite-dimensional , embedding it as a subalgebra of \mathfrak{gl}(n, K) for some n. These results underscore the realizability of simple in matrix , facilitating their study through linear algebra. Over the complex numbers \mathbb{C}, all finite-dimensional simple Lie algebras are completely classified into four infinite families of classical types—A_n (for n \geq 1), B_n (for n \geq 2), C_n (for n \geq 3), and D_n (for n \geq 4)—along with five exceptional types: G_2, F_4, E_6, E_7, and E_8. This classification, originally due to Killing and Cartan, relies on the structure of root systems and the properties of the Killing form, providing a complete list up to isomorphism without further infinite families.

Semisimple Lie Groups

A connected G is semisimple if its \mathfrak{g} is semisimple, meaning \mathfrak{g} has no nonzero solvable ideals. This condition ensures that G has no nontrivial connected normal solvable subgroups, distinguishing semisimple Lie groups from more general reductive ones that may include abelian factors. As detailed in the section on , a semisimple decomposes as a of . The structure theorem for semisimple Lie groups states that every connected semisimple group is a finite cover of a of simple groups. For compact connected semisimple groups, this implies that G is a of a product K_1 \times \cdots \times K_r of connected compact simple groups by a finite central . In the simply connected case, the group is precisely a of simply connected simple groups. Semisimple Lie groups may possess nontrivial centers, which arise from the covering structure and can be finite even for noncompact examples. In contrast, the adjoint form of a simple Lie group has trivial , as the is faithful. The universal of a semisimple Lie group thus reveals the full central structure, often compact with a finite . In the Levi decomposition of the Lie algebra of a semisimple , the is trivial, and the coincides with the itself. For instance, the \mathrm{[SL](/page/SL)}(n, \mathbb{C}) for n \geq 2 is a simple , hence semisimple, with \mathfrak{[sl](/page/SL)}(n, \mathbb{C}) simple. However, a such as \mathrm{[SL](/page/SL)}(m, \mathbb{C}) \times \mathrm{[SL](/page/SL)}(n, \mathbb{C}) for m, n \geq 2 is semisimple but not simple, as it contains proper subgroups isomorphic to each .

Relationship to Root Systems

The structure of a simple Lie algebra \mathfrak{g} over \mathbb{C} is intimately connected to its through the root space decomposition. For a \mathfrak{h} \subset \mathfrak{g}, the Lie algebra decomposes as \mathfrak{g} = \mathfrak{h} \oplus \bigoplus_{\alpha \in \Delta} \mathfrak{g}_\alpha, where \Delta is the consisting of the nonzero weights of the of \mathfrak{h} on \mathfrak{g}, and \mathfrak{g}_\alpha = \{ x \in \mathfrak{g} \mid \operatorname{ad}_h x = \alpha(h) x \ \forall h \in \mathfrak{h} \} is the root space for each root \alpha \in \Delta. This decomposition highlights how the \Delta encodes the semisimple nature of \mathfrak{g}, with each \dim \mathfrak{g}_\alpha = 1 for finite-dimensional simple Lie algebras. A choice of positive roots \Delta^+ within \Delta yields a basis of simple roots \Pi = \{\alpha_1, \dots, \alpha_r\}, where r = \dim \mathfrak{h} is the , and every positive root is a non-negative integer combination of elements from \Pi. The W of the is the generated by reflections s_\alpha: v \mapsto v - 2 \frac{(\alpha, v)}{(\alpha, \alpha)} \alpha across the hyperplanes perpendicular to roots \alpha \in \Delta, acting faithfully on the span of \Delta. For simple Lie algebras, the \Delta is irreducible, meaning it cannot be decomposed into orthogonal subsystems, which ensures the algebra has no nontrivial ideals. These root systems are classified up to by their associated Coxeter-Dynkin diagrams, which capture the intersection angles and lengths of simple roots. Central to the root system are the coroots \alpha^\vee = \frac{2\alpha}{(\alpha, \alpha)}, where (\cdot, \cdot) is the invariant bilinear form on the dual space. The Cartan matrix A = (a_{ij}) is defined by a_{ij} = (\alpha_i, \alpha_j^\vee) for i, j = 1, \dots, r, encoding the structure constants of the Lie algebra via the relations [\mathfrak{g}_{\alpha_i}, \mathfrak{g}_{\alpha_j}] \subseteq \mathfrak{g}_{\alpha_i + \alpha_j]. This matrix is crucial for determining the Dynkin diagram labels and distinguishing irreducible root systems. The relationship extends to the corresponding simple Lie group G via the exponential map \exp: \mathfrak{g} \to G, which associates elements of root spaces to one-parameter subgroups. Specifically, for X \in \mathfrak{g}_\alpha with \alpha \in \Delta, the curve t \mapsto \exp(tX) generates a one-parameter subgroup in G whose Lie algebra element reflects the root structure, facilitating the study of maximal tori and conjugacy classes in G.

Classification Overview

Historical Development

The foundations of the theory of simple Lie groups were laid in the late 19th century by , who developed the concept of continuous transformation groups as a tool for solving differential equations, introducing the infinitesimal approach that would later underpin Lie algebras. Lie's work, spanning the and , emphasized the geometric and analytical properties of these groups, establishing them as central to understanding symmetries in continuous settings. A pivotal advancement came in 1888–1890 when classified the simple Lie algebras over the complex numbers, identifying the basic structures through the introduction of Cartan subalgebras and matrices, though his proofs contained some gaps. refined and completed this classification in his 1894 doctoral thesis, providing rigorous constructions for semisimple algebras over the complex field, and extended it to real forms between 1894 and 1913, addressing the structure of non-compact cases. In 1913–1914, Cartan proved the uniqueness and completeness of this classification for finite-dimensional simple Lie algebras, solidifying the framework for both complex and real settings. Hermann Weyl advanced the theory in 1925 through his work on representation theory of semisimple Lie groups, integrating matrix representations to reveal deeper structural insights and influencing quantum mechanics applications. In the 1950s, Claude Chevalley developed the theory of algebraic groups over arbitrary fields, classifying semisimple groups and introducing Chevalley groups as finite analogues of Lie groups, which extended the classification to positive characteristic settings. Concurrently, Harish-Chandra's contributions in the 1950s focused on representations of real semisimple Lie groups, establishing key results on unitary representations and harmonic analysis that completed the understanding of non-compact real forms. By the mid-1950s, the classification of finite-dimensional simple Lie groups and algebras over and numbers was exhaustive, with Chevalley's work marking the culmination of efforts to cover all cases, and no new types have been discovered since. This completeness has provided a stable foundation for subsequent developments in and .

Cartan-Dynkin Diagrams

Cartan-Dynkin diagrams provide a visual encoding of the simple systems associated with simple Lie algebras, facilitating their up to . Each diagram consists of nodes representing the simple of the , with connecting pairs of nodes to indicate the inner product between the corresponding . Specifically, the absence of an denotes orthogonal (inner product zero), a single represents with inner product -1 (corresponding to an of 120°), a double indicates an inner product of -1 with one twice as long as the other ( 135°), and a triple signifies an inner product of -3/2 with the length ratio 3:1 ( 150°). Arrows on multiple point from the longer to the shorter one, ensuring the diagram captures the of the precisely. These diagrams are closely related to but distinct from Coxeter diagrams. Dynkin diagrams emphasize the connections via coroots, incorporating directional arrows to account for root length differences and directly encoding the entries A_{ij} = 2 \langle \alpha_i, \alpha_j \rangle / \langle \alpha_i, \alpha_i \rangle, where \alpha_i are roots and \langle \cdot, \cdot \rangle is the inner product. In , Coxeter diagrams describe the relations in the generated by reflections across the root hyperplanes, using multiple edges (without arrows) to denote the order of the product of two reflections, such as 3 for a single edge (120° angle), 4 for double, and 6 for triple. This distinction highlights how Dynkin diagrams are tailored to structure, while Coxeter diagrams generalize to finite reflection groups. A key property is that each connected Cartan-Dynkin corresponds to a unique irreducible , meaning the simple roots cannot be decomposed into orthogonal subsets, which ensures the associated is simple rather than semisimple with multiple factors. The connectivity of the thus guarantees indecomposability, aligning with the of finite-dimensional simple over the numbers. For example, the for the series A_n (corresponding to \mathfrak{sl}(n+1, \mathbb{C})) is a linear chain of n nodes connected by single edges, reflecting the equal-length roots in a . In contrast, the exceptional for E_8 features eight nodes: a chain of seven nodes with single edges, and a branch from the third node in the chain to the eighth node via another single edge, capturing the intricate structure of its 248-dimensional . The utility of Cartan-Dynkin diagrams lies in their ability to fully determine the from the graph structure, which in turn specifies the up to , as the matrix governs the and the Chevalley basis. By reconstructing the matrix—diagonal entries 2, off-diagonals -1, -2, or -3 based on edge multiplicity and direction—one can generate the entire and thus the , providing a compact tool for studying representations and group structures. This encoding has proven essential in applications from to .

Types of Simple Lie Groups

Simple Lie groups are classified according to the Killing-Cartan classification of their underlying Lie algebras, which divides them into classical and exceptional types based on the structure of their root systems. The classical series consist of four infinite families: the A_n series (n ≥ 1), corresponding to special linear groups (n+1, ℂ) over the complex numbers or their real forms like (n+1, ℝ); the B_n series (n ≥ 2), associated with odd orthogonal groups SO(2n+1, ℂ) or real forms such as SO(2n+1, ℝ); the C_n series (n ≥ 3), linked to groups Sp(2n, ℂ) or real forms like Sp(2n, ℝ); and the D_n series (n ≥ 4), corresponding to even orthogonal groups SO(2n, ℂ) or real forms including SO(n, n). These families capture the majority of simple Lie groups and arise from matrix groups preserving bilinear or sesquilinear forms. The exceptional series comprises five finite-dimensional types: G_2, F_4, E_6, E_7, and E_8, each with corresponding complex Lie groups and multiple real forms that do not fit into the classical matrix group patterns. For instance, G_2 and F_4 have real forms including split and compact variants, while E_6, E_7, and E_8 exhibit additional non-compact real forms beyond the split and compact ones. These exceptional groups are distinguished by their unique root systems and play key roles in areas like and . Complex simple Lie groups are those whose Lie algebras are complex simple, meaning they are defined over ℂ and have no non-trivial complex ideals; their simply connected covers are complex manifolds. In contrast, real simple Lie groups have real Lie algebras that are simple over ℝ, which are real forms of the above and include compact forms (with negative definite ), split forms (with a maximal split torus, where the is fully real and diagonalizable over ℝ), and other non-compact forms. Split forms are particularly significant as they correspond to real points of algebraic groups defined over ℝ, maximizing the dimension of the maximal split torus. The yields four infinite classical families (A_n, B_n, C_n, D_n) and five exceptional types (G_2, F_4, E_6, E_7, E_8) for the cases, with each admitting finitely many real forms; notably, the exceptional types collectively possess 14 distinct real forms across their ranks. This structure provides a complete typology, previewing deeper analyses of specific forms like compact real simple Lie groups.

Detailed Classification

Compact Simple Lie Groups

Compact simple Lie groups are connected, non-abelian Lie groups with no nontrivial connected subgroups and a compact topology as manifolds. Their associated Lie algebras are real forms of complex simple Lie algebras where the Killing form is negative definite, ensuring the group's compactness. Representative examples include the special unitary groups SU(n) for n ≥ 2, the compact symplectic groups Sp(n) for n ≥ 1, the spin groups (n) for n ≥ 3, and the exceptional groups G₂, F₄, E₆, E₇, and E₈. A defining property of compact simple Lie groups is that all finite-dimensional representations over the complex numbers are completely reducible, meaning they decompose into a of irreducible representations without invariant subspaces. This follows from the existence of a bi-invariant , which is finite and unique up to scalar multiple, allowing the use of averaging operators to project onto invariant subspaces. Additionally, the negative definiteness of the Killing form on the implies that the group admits an Ad-invariant inner product, facilitating the study of its geometry and representations. The classification of compact simple Lie groups parallels that of complex simple Lie algebras, as each such group arises as the connected component of the maximal compact of the adjoint form of a semisimple Lie group, or equivalently, as the compact real form of a simple Lie algebra. This classification, originally established by Killing and refined by Cartan, uses root systems and their Dynkin diagrams to enumerate four infinite families—Aₙ corresponding to SU(n+1), Bₙ to (2n+1), Cₙ to Sp(n), and Dₙ to (2n)—along with five exceptional cases: G₂, F₄, E₆, E₇, and E₈. All compact simple Lie groups are covered by these types, with the simply laced ones (Aₙ, Dₙ, E series) having roots of equal length. A fundamental structural result is that every compact simple Lie group G is isomorphic to a quotient of its simply connected cover Ĝ by a finite central subgroup Z, where Z is a discrete subgroup of the center of Ĝ, ensuring that the fundamental group of G is finite. This isogeny structure preserves the Lie algebra and root system while accounting for the group's topology. The representation theory of compact simple Lie groups is governed by the Peter-Weyl theorem, which states that the matrix coefficients of the irreducible unitary representations form an for the L²(G) with respect to the normalized . Consequently, every on G can be uniformly approximated by finite linear combinations of these coefficients. Furthermore, the irreducible finite-dimensional representations are in bijective correspondence with dominant integral weights in the fundamental Weyl chamber of the dual of the , parameterized by the orbits.

Complex Simple Lie Groups

Complex simple Lie groups are connected, finite-dimensional Lie groups over the complex numbers whose associated Lie algebras are simple, meaning they are non-abelian and contain no nontrivial ideals. These groups have no nontrivial connected subgroups and are semisimple. Representative examples include the \mathrm{SL}(n, \mathbb{C}) for n \geq 2, corresponding to the type A_{n-1} , and the \mathrm{Sp}(2n, \mathbb{C}) for n \geq 1, of type C_n. As complex manifolds, these groups are Stein spaces, characterized by being holomorphically convex with no non-constant bounded holomorphic functions, which facilitates the study of their global analytic properties. Their finite-dimensional representations are holomorphic, meaning they extend to analytic homomorphisms into matrix groups like \mathrm{GL}(k, \mathbb{C}), and are completely reducible due to the semisimplicity of the underlying algebras. The structure of these groups is intimately tied to their root systems, allowing decomposition into factors resembling \mathrm{SL}(2, \mathbb{C}) connected via Serre relations defined by Dynkin diagrams. A key aspect of complex simple Lie groups is their relation to real forms, where each admits a unique compact real form that acts as its maximal compact subgroup and determines the group's type. For example, \mathrm{SU}(n) is the compact real form of \mathrm{SL}(n, \mathbb{C}). This compact subgroup embeds as a real , preserving the complex structure while providing a finite-dimensional model for topological invariants. The classification of complex simple Lie groups mirrors that of their Lie algebras, achieved through the enumeration of irreducible root systems over \mathbb{C}, yielding four infinite classical series (A_n, B_n, C_n, D_n) and five exceptional types (E_6, E_7, E_8, F_4, G_2). This classification, established by Killing and Cartan, relies on the non-degeneracy of the Killing form and the structure of Weyl chambers, ensuring all such groups are captured without infinite-dimensional extensions in the finite case. Complex simple Lie groups are semisimple with no non-trivial compact factors in their complex structure, though their study often leverages the real compact forms for representation theory and topology. Their centers are finite, detected via the quotient of weight and root lattices, varying by type—for instance, cyclic of order n+1 for A_n.

Real Non-Compact Simple Lie Groups

Real non-compact simple Lie groups are the connected Lie groups whose Lie algebras are non-compact real forms of complex simple Lie algebras. A real form of a complex simple Lie algebra \mathfrak{g}_\mathbb{C} is a real subalgebra \mathfrak{g} such that \mathfrak{g}_\mathbb{C} = \mathfrak{g} \otimes_\mathbb{R} \mathbb{C}, and \mathfrak{g} is simple if it has no nontrivial ideals. These real forms are classified up to isomorphism by Élie Cartan, who determined all simple real Lie algebras in 1914 by studying the possible real structures compatible with the complex structure. The classification distinguishes the compact real form, where the Killing form is negative definite, from non-compact ones, where the Killing form has both positive and negative eigenvalues, leading to indefinite signature. Satake diagrams provide a compact graphical encoding of this classification: starting from the Dynkin diagram of the complex type, vertices corresponding to compact imaginary roots are painted black, while pairs of non-compact imaginary roots are connected by arrows, and real roots remain white unpainted vertices. A real form is non-compact if its Satake diagram has at least one white vertex or arrow, as opposed to the all-black diagram of the compact form. Among the non-compact real forms, the split real forms are those admitting a that is \mathbb{R}-, meaning it is the of one-dimensional ad-diagonalizable subspaces over \mathbb{R}, or equivalently, all are real with respect to some ordering. In Satake diagrams, split forms appear as the original unpainted with no arrows or black vertices, maximizing the rank of the split . Representative examples include \mathfrak{sl}(n,\mathbb{R}) for type A_{n-1} and \mathfrak{so}(p,q) with p = q + 1 or p = q for types B_n and D_n, where the split form achieves the highest for the maximal abelian subspace in the symmetric complement \mathfrak{p} of the Cartan decomposition \mathfrak{g} = \mathfrak{k} \oplus \mathfrak{p}, with \mathfrak{k} the maximal compact . Split forms play a central role in the and geometry of these groups, as their root systems over \mathbb{R} mirror the complex case most closely. Other non-compact real forms include quasi-split forms, where the derived algebra of the centralizer of a Borel is the nilradical, and more specialized types such as quaternionic forms, which admit a quaternionic and correspond to Satake diagrams with specific arrow patterns indicating non-real compact . For instance, in type A_1, the unique non-compact form is \mathfrak{sl}(2,\mathbb{R}), the form itself. Quaternionic examples arise in types like A_{n-1} (for even dimensions) as \mathfrak{su}^*(2n), isomorphic to \mathfrak{sl}(n,\mathbb{H}), and in type C_n beyond the split \mathfrak{sp}(2n,\mathbb{R}). Exceptional non-compact forms, not fitting classical patterns, occur in the exceptional types, such as the \mathfrak{g}_{2(2)} for G_2 or \mathfrak{f}_{4(-20)} for F_4, distinguished by their Satake diagrams with mixed black vertices and arrows. These forms are inner or outer depending on whether the diagram respects the of the Dynkin . The number of distinct real forms (up to ) for a given type varies: for type A_n (n \geq 1), there are \left\lfloor \frac{(n+1)^2}{4} \right\rfloor + 1 total forms, with all but one being non-compact for n \geq 1; for B_n and D_n, the count is roughly n+1; for C_n, it is n+1; and exceptional types have fewer, typically 1 to 4 non-compact forms each (e.g., 2 for E_8, 4 for E_6). This multiplicity arises from the possible signatures in the Cartan decomposition, constrained by the . A fundamental structural theorem for the associated connected non-compact Lie groups G is the G = K A N, where K is the maximal compact , A is a maximal \mathbb{R}-split , and N is a maximal unipotent , with the product being direct at the level \mathfrak{g} = \mathfrak{k} \oplus \mathfrak{a} \oplus \mathfrak{n} and unique for each g \in G. This decomposition, established by , facilitates and the study of representations on these groups.

Examples and Lists

Low-Dimensional Simple Lie Groups

Simple Lie groups of dimension 3 are the smallest non-abelian examples and illustrate the basic structure of the classification. The unique compact simple Lie group in this dimension is \mathrm{SU}(2), which is isomorphic to the symplectic group \mathrm{Sp}(1) over the quaternions and to the spin group \mathrm{Spin}(3). This group is simply connected and double covers the orthogonal group \mathrm{SO}(3), with Lie algebra \mathfrak{su}(2). The corresponding non-compact simple Lie group is \mathrm{SL}(2, \mathbb{R}), which is isomorphic to \mathrm{SU}(1,1) and has Lie algebra \mathfrak{sl}(2, \mathbb{R}); it serves as the split real form of type A_1. These isomorphisms highlight the duality between compact and non-compact forms in low dimensions. In dimension 6, there is a unique simple Lie group up to , namely \mathrm{SL}(2, \mathbb{C}), the over the complex numbers viewed as a real manifold. Its Lie algebra \mathfrak{sl}(2, \mathbb{C}), considered over \mathbb{R}, is simple and isomorphic to \mathfrak{so}(3,1), the Lie algebra of the \mathrm{SO}^+(3,1). This example represents the complex real form of type A_1, where the group structure over \mathbb{C} yields a 6-dimensional real without nontrivial connected normal subgroups. The next low-dimensional cases occur in dimension 8, corresponding to the root system A_2. The compact simple Lie group is \mathrm{SU}(3), with Lie algebra \mathfrak{su}(3) of dimension 8, which is simply connected and covers the projective group \mathrm{PSU}(3). The split real form is \mathrm{SL}(3, \mathbb{R}), with Lie algebra \mathfrak{sl}(3, \mathbb{R}), while the non-split non-compact form is \mathrm{SU}(2,1), with Lie algebra \mathfrak{su}(2,1). These three real forms exhaust the possibilities for type A_2, distinguished by their signatures and the restricted root systems. In 10, the simple Lie groups arise from the isomorphic types B_2 and C_2. The compact form is realized by \mathrm{SO}(5), which has \mathfrak{so}(5), and is locally isomorphic to \mathrm{Sp}(2), the compact over the quaternions, with simply connected cover \mathrm{Spin}(5) \cong \mathrm{Sp}(2). The non-compact forms include the split real form \mathrm{SO}(4,1) or \mathrm{Sp}(4, \mathbb{R}), and the quasi-split form \mathrm{SO}(3,2). The B_2 \cong C_2 as algebras leads to shared real forms in this dimension. Up to dimension 10, these exhaust the simple Lie groups: the A_1 forms in dimensions 3 and 6, the A_2 forms in dimension 8, and the B_2 \cong C_2 forms in dimension 10. No other simple Lie groups exist in dimensions 4, 5, 7, or 9, reflecting the constraints of the Cartan-Killing on possible systems and real forms.

Classical Simple Lie Groups

The classical simple Lie groups comprise four infinite families, denoted A_n, B_n, C_n, and D_n for (with some low-dimensional exceptions and adjustments), corresponding to the special linear, odd orthogonal, , and even orthogonal groups, respectively. These groups arise as matrix groups preserving specific bilinear forms and form the backbone of the of simple Lie groups over the complex numbers, with various real forms providing non-compact realizations. Their Lie algebras are simple and share systems that distinguish them from the exceptional types. The A_n series is associated with the special linear group SL(n+1, \mathbb{C}), the group of (n+1) \times (n+1) complex matrices with determinant 1, modulo its center to yield the adjoint form; its Lie algebra \mathfrak{sl}(n+1, \mathbb{C}) consists of trace-zero matrices and has dimension n^2 + 2n. The fundamental representation acts on \mathbb{C}^{n+1} by matrix multiplication. Real forms include the compact SU(n), preserving the Hermitian inner product; the split form SL(n, \mathbb{R}), preserving the real determinant; and the indefinite SU(p, q) with p + q = n+1, preserving a Hermitian form of signature (p, q). The B_n series corresponds to the odd orthogonal group SO(2n+1, \mathbb{C}), preserving a non-degenerate symmetric bilinear form on \mathbb{C}^{2n+1}; its Lie algebra \mathfrak{so}(2n+1, \mathbb{C}) has dimension n(2n+1) and consists of skew-symmetric matrices with respect to this form. The fundamental representation is the vector representation on \mathbb{C}^{2n+1}. The compact real form is SO(2n+1), while a non-compact form is SO(n, n+1), preserving a quadratic form of signature (n, n+1). The simply connected cover is Spin(2n+1). For the C_n series, the complex group is Sp(n, \mathbb{C}) \cong Sp(2n, \mathbb{C}), the group of 2n \times 2n complex matrices preserving a skew-symmetric bilinear form on \mathbb{C}^{2n}; the Lie algebra \mathfrak{sp}(n, \mathbb{C}) has dimension n(2n+1). The fundamental representation acts on \mathbb{C}^{2n} as the defining symplectic action. The compact real form is Sp(n), the compact symplectic group (also denoted USp(2n)), while non-compact forms include Sp(2n, \mathbb{R}) and Sp(p, q) preserving a quaternionic Hermitian form of appropriate signature. The D_n series is realized by SO(2n, \mathbb{C}), preserving a symmetric bilinear form on \mathbb{C}^{2n}; its Lie algebra \mathfrak{so}(2n, \mathbb{C}) has dimension n(2n-1). The fundamental representation is the vector action on \mathbb{C}^{2n}. Compact real forms include SO(2n) and its double cover Spin(2n), while non-compact forms are SO(p, q) with p + q = 2n. For low dimensions, exceptional isomorphisms occur: SU(4) \cong Spin(6) and Sp(2) \cong Spin(5).

Exceptional Simple Lie Groups

The exceptional simple Lie groups form a finite family of five types, denoted G_2, F_4, E_6, E_7, and E_8, which do not fit into the infinite classical series of types A_n, B_n, C_n, and D_n. These groups arise uniquely from the of simple Lie algebras over the complex numbers and exhibit intricate structures tied to division algebras beyond the reals, complexes, and quaternions. They form a chain of inclusions G_2 \subset F_4 \subset E_6 \subset E_7 \subset E_8, reflecting their nested root systems, and their dimensions are 14, 52, 78, 133, and 248, respectively. Unlike the classical groups, the exceptional ones lack uniform matrix realizations but appear in constructions involving and Jordan algebras, with applications in geometry and physics. The smallest exceptional simple Lie group is G_2, of 14 and 2, whose compact real form is the of the , preserving their multiplication table. Over the complexes, G_2 has a unique real form up to that is compact, but there are two distinct real forms: the compact G_2 and the non-compact split form G_{2(2)}, with maximal compact subgroup \mathrm{SL}(2,\mathbb{R}) \times \mathrm{SL}(2,\mathbb{R}). The group G_2 is not simply laced, featuring short and long roots in its . Next is F_4, of dimension 52 and rank 4, whose compact real form admits realizations as the of the exceptional of 3×3 Hermitian octonionic matrices, denoted \mathfrak{h}_3(\mathbb{O}). This algebra encodes a 27-dimensional structure, and F_4 preserves both the Jordan product and the real part of the trace form on \mathfrak{h}_3(\mathbb{O}). There are three real forms: the compact F_4, the non-compact F_{4(-20)} with signature related to Hermitian symmetric spaces, and F_{4(-52)}, the split form with maximal compact subgroup \mathrm{Spin}(9). Like G_2, F_4 has a non-simply laced . The group E_6, of dimension 78 and rank 6, features a fundamental 27-dimensional representation exhibiting triality, an outer automorphism of order 3 that permutes three equivalent embeddings related to octonionic structures and links to \mathrm{Spin}(8) representations. Its compact real form contains maximal subgroups such as \mathrm{SU}(3)^3, arising from the extended Dynkin diagram where three nodes correspond to the three \mathrm{SU}(3) factors. There are four real forms: the compact E_6, the split E_{6(6)} with maximal compact \mathrm{Sp}(4), the quasi-split E_{6(2)} with maximal compact \mathrm{SU}(6) \times \mathrm{SU}(2), and E_{6(-14)} with maximal compact \mathrm{Spin}(10) \times \mathrm{SO}(2); a further form E_{6(-26)} has maximal compact F_4. E_6 is simply laced, with all roots of equal length.) E_7, of dimension 133 and rank 7, possesses a fundamental representation of dimension 56 and includes E_6 \times \mathrm{SU}(2) as a maximal . Among its three real forms—the compact E_7, the E_{7(7)}, and E_{7(-25)}—the latter is Hermitian, corresponding to a symmetric space of non-compact type with maximal compact E_6 \times \mathrm{SO}(2). This Hermitian form plays a role in bounded symmetric domains and Freudenthal geometries. E_7 is also simply laced. Finally, E_8, of 248 and 8, is the largest exceptional simple Lie group, with its compact real form serving as the of the unique 248-dimensional Freudenthal triple system over the . It contains the full chain of lower exceptional groups as subalgebras and has two real forms: the compact E_8 and the split non-compact E_{8(8)}, with no further non-compact forms due to the algebra's rigidity. E_8 is simply laced and lacks finite-dimensional representations beyond the . Key structural features of the exceptional groups include subgroup embeddings beyond the inclusion chain, such as \mathrm{SU}(3)^3 \subset E_6, and the , a 3×3 array constructed from pairs of normed division algebras (reals, complexes, quaternions, ) that systematically generates the exceptional Lie algebras in the off-diagonal entries via derivations and structurable algebra operations. This square highlights the exceptional groups' ties to algebraic structures over non-associative algebras.

Special Properties and Structures

Simply Laced Simple Lie Groups

Simply laced simple Lie groups are those whose associated root systems consist of roots all of the same length. This property distinguishes them within the classification of simple Lie groups, corresponding precisely to the of : the for n ≥ 1, the for n ≥ 4, and the exceptional cases , , . In these diagrams, all edges are single bonds with no multiple connections, reflecting the uniform root lengths and symmetric . A key property of simply laced simple Lie groups is that their dual Coxeter number h^∨ coincides with the Coxeter number h, due to the absence of short and long roots. For the A_n series, which has rank n, this number equals n + 1. Their affine extensions form the simply laced affine Lie algebras, which extend the finite-dimensional structure while preserving the uniform root lengths in the extended root systems. In the representation theory of simply laced simple Lie groups, the roots are conventionally normalized using the Killing form such that the squared length (α, α) = 2 for all roots α, corresponding to a Euclidean length of √2; this eliminates any distinction between short and long roots that arises in other cases. Representative examples include the special unitary groups SU(n+1), associated with the A_n series; the special orthogonal groups SO(2n) for n ≥ 4, associated with D_n; and the exceptional simple Lie groups of types E_6, E_7, and E_8. In contrast, the simple Lie groups of types B_n, C_n, F_4, and G_2 are not simply laced, as their root systems include roots of two distinct lengths, leading to asymmetric Cartan matrices and multiple bonds in their Dynkin diagrams.

Hermitian Symmetric Spaces

Hermitian symmetric spaces are a special class of Riemannian symmetric spaces equipped with an invariant complex structure compatible with the metric. Specifically, for a simple non-compact Lie group G, these spaces are realized as quotients G/K, where K is a maximal compact subgroup of G, and the tangent space at the base point admits a complex structure J such that J^2 = -I and the metric is preserved under J. This structure ensures that the space is Kähler, with the Kähler form defined by \omega(X, Y) = g(X, JY), where g is the invariant metric. The classification of irreducible Hermitian symmetric spaces of non-compact type consists of four classical series and two exceptional cases, corresponding to specific real forms of the complex simple Lie algebras: SU(p,q)/S(U(p)×U(q)) for p,q ≥1 (type AIII, associated to A_{p+q-1}), SO*(2n)/U(n) for n ≥2 (type DIII, D_n), Sp(2n,ℝ)/U(n) for n ≥1 (type CI, C_n), SO(2,n)/SO(2)×SO(n) for n ≥2 (type IV_n, associated to real forms of B_n or D_{(n/2)+1}), E_6^{-14}/(Spin(10)×U(1)) (type EIII, E_6), and E_7^{-25}/(E_6×U(1)) (type EVII, E_7). Classical examples include the groups SU(p,q)/S(U(p) × U(q)) for p,q ≥1, which realize the type AIII series. These spaces are biholomorphic to bounded symmetric domains in complex vector spaces, and their classification follows from the representation theory of the restricted root system, where the center of K acts non-trivially on the tangent space. Key properties include the natural Kähler metric induced by the complex structure, which is invariant under the , and the existence of a Bergman kernel function k_D(z, w) that defines the metric on the bounded realization. The embedding provides a holomorphic realization of the space as a bounded in the of the Cartan , mapping the symmetric space into a diffeomorphic to a or higher-dimensional analogue. These embeddings highlight the geometric duality between non-compact Hermitian spaces and their compact duals. Representative examples are the complex hyperbolic space \mathbb{CH}^n, realized as SO(2,n)/ (SO(2) × SO(n)) or SU(n,1)/U(n), which corresponds to the acting on with complex structure, and the upper half-plane, given by Sp(2n,\mathbb{R})/U(n), parametrizing symmetric complex matrices with positive definite imaginary part. In the first case, the space is the unit ball in \mathbb{C}^n, while the second is a tube domain over the cone. The group G acts transitively on the G/K, with the stabilizer at a base point being the maximal compact subgroup K. Furthermore, the G = K A N identifies minimal parabolic subgroups whose unipotent radicals N stabilize points on the Furstenberg boundary, facilitating the study of flows and representations.

Applications in Physics

Simple Lie groups underpin the symmetries of the fundamental interactions in . In (QCD), the strong nuclear force is mediated by gluons transforming under the of the simple Lie group SU(3), which governs the color symmetry among quarks. This non-Abelian structure leads to at high energies and confinement at low energies, explaining the observed properties of hadrons. The electroweak sector of the relies on the semisimple gauge group SU(2) × U(1), where SU(2) provides the simple Lie group factor for symmetry, unifying the weak and electromagnetic forces through via the . Grand unified theories (GUTs) extend this framework by embedding the gauge groups into larger simple Lie groups to unify the strong, weak, and electromagnetic interactions at high energies. The SU(5) model, proposed by Georgi and Glashow, unifies SU(3) × SU(2) × U(1) within SU(5), predicting and gauge coupling unification around 10^{15} GeV, though experimental bounds have constrained its minimal form. Similarly, the SO(10) GUT incorporates all fermions of one generation into the 16-dimensional spinor representation, naturally including right-handed s and enabling mechanisms for neutrino masses. Exceptional simple Lie groups like E_6 appear in GUT models that further unify matter representations, such as the 27-dimensional fundamental, accommodating three generations while addressing flavor symmetries and lepton-quark unification. In , simple Lie groups structure the gauge sectors of compactifications. The E_8 × E_8 embeds the gauge group within one or both E_8 factors, providing a framework for grand unification and cancellation in ten dimensions. Compactifications on Calabi-Yau manifolds utilize the of simple Lie algebras (corresponding to SU(n), SO(2n), and exceptional groups) to resolve singularities, yielding enhanced gauge symmetries in the low-energy effective theory. Conformal field theories (CFTs) on worldsheets or in two-dimensional critical systems incorporate affine Kac-Moody algebras based on simple Lie groups, extending the through the Sugawara to generate stress-energy tensors and ensure conformal invariance. These Wess-Zumino-Witten models describe symmetries in condensed matter systems and string backgrounds. Recent applications post-2000 highlight simple Lie groups in emerging fields. In , Clifford groups—derived from the Spin groups via Clifford algebras—form the basis for stabilizer codes and fault-tolerant universal gates when augmented by non-Clifford operations like T-gates. Integrable systems with G_2 symmetries, such as systems on moduli spaces, model exceptional gauge theories and provide exact solutions in supersymmetric quantum field theories.

References

  1. [1]
    [PDF] notes on simple lie algebras and lie groups. math 261a
    Definition 0.3. A simple Lie group is a connected non-abelian Lie group G which does not have nontrivial connected (analytic) normal. Lie subgroups.
  2. [2]
    [PDF] Introduction to Lie Groups and Lie Algebras Alexander Kirillov, Jr.
    Lie Groups: Basic. Definitions. 2.1. Lie groups, subgroups, and cosets. Definition 2.1. A Lie group is a set G with two structures: G is a group and G is a ...
  3. [3]
    [PDF] the classification of simple complex lie algebras - UChicago Math
    Aug 24, 2012 · Definition 2.1. A Lie group G is a smooth manifold with a group operation such that multiplication m : G × G → G and inversion i : G → ...
  4. [4]
    [PDF] lie groups, lie algebras, and applications in physics - UChicago Math
    Sep 17, 2015 · This paper introduces basic concepts from representation theory,. Lie group, Lie algebra, and topology and their applications in physics, par-.
  5. [5]
    [PDF] Lie Groups. Representation Theory and Symmetric Spaces
    Definition 1.1 A Lie group G is an abstract group and a smooth n- dimensional manifold so that multiplication G × G → G: (a, b) → ab and inverse G → G: a → a−1 ...
  6. [6]
    simple Lie group in nLab
    ### Summary of Simple Lie Group Definition and Key Theorems
  7. [7]
    [PDF] Introduction to Lie Groups and Lie Algebras Alexander Kirillov, Jr.
    In the study of root systems of simple Lie algebras, we will frequently use the following lattices. ... Warner, Harmonic analysis on semi-simple Lie groups. I, II ...
  8. [8]
  9. [9]
    [PDF] Introduction to Lie Algebras and Representation Theory
    This book is designed to introduce the reader to the theory of semisimple. Lie algebras over an algebraically closed field of characteristic 0, with emphasis on ...
  10. [10]
    [PDF] Complex Semisimple Lie Algebras - Pierre Clare
    For n ~ 1, An= sl(n + 1) is the Lie algebra of the special linear group in n + 1 variables, SL(n + 1). For n ~ 2, Bn = so(2n + 1) is the Lie algebra of the ...Missing: A_n B_n C_n D_n
  11. [11]
    [PDF] LIE GROUPS AND LIE ALGEBRAS
    Moreover, all Lie algebras su(p, q) are semisimple. The compact real form is obtained by setting p = n, q = 0 (or p = 0 and q = n), resulting in su(n). The ...
  12. [12]
    [PDF] LECTURE 9. Definition 1.1. We say that a Lie algebra is simple if it is ...
    SIMPLE AND SEMISIMPLE LIE ALGEBRAS. Definition 1.1. We say that a Lie algebra is simple if it is non-Abelian and has no nontrivial proper ideal.
  13. [13]
    [PDF] 5. Semisimple Lie algebras and the Killing form - Brandeis
    (3) The Killing form of L is nondegenerate. (4) L is a direct sum of simple ideals. 5.1. Definition. First we observe that the sum of two solvable ideals ...
  14. [14]
    Appendix B. Lie's Third Theorem - Project Euclid
    Jan 1, 2023 · This result sharpens Lie's Third Theorem, saying that every real Lie algebra is the Lie algebra of an analytic group of matrices. The Campbell- ...
  15. [15]
    [PDF] ADO'S THEOREM 1. Introduction Let L be a finite-dimensional Lie ...
    The purpose of this expository note is to give a proof of Ado's Theorem, which states that every finite-dimensional Lie algebra over the field of characteristic ...
  16. [16]
    [PDF] Lectures on Lie Groups Dragan Milicic - Math
    simple Lie algebras of dimension ≤ 2. On the other hand, assume that G = SU ... compact simple Lie groups. 1.9. Fundamental group of a connected ...
  17. [17]
    [PDF] Lie Algebras, Algebraic Groups, and Lie Groups - James Milne
    May 5, 2013 · In Chapter III we show that all connected complex semisimple Lie groups are algebraic groups, and that all connected real semisimple. Lie ...
  18. [18]
    [PDF] Lecture 15 — Classical (Semi) Simple Lie Algebras and Root Systems
    Oct 28, 2010 · a) Using the root space decomposition, prove that soN (F) is semisimple if N ≥ 3. b) Show soN (F) is simple if N = 3 or N ≥ 5 by showing ...
  19. [19]
    [PDF] 47 Construction of Lie algebras from a root lat- tice
    The root space decomposition of a Lie algebra suggests the following construc- tion of a Lie algebra from its root system. Take the direct sum of the (dual.
  20. [20]
    [PDF] Lecture 21 — The Weyl Group of a Root System
    Nov 23, 2010 · The group W is called the Weyl group of the root system (V,∆) (and of the corresponding semisimple lie algebra g). ... simple roots (r = dimV ).
  21. [21]
    [PDF] Classification of root systems
    Sep 8, 2017 · The root system is called simple if R is a single equivalence class. We will recall in Section 2 the connection with semisimple Lie algebras.
  22. [22]
    [PDF] Reference sheet for classical roots systems - UC Berkeley math
    Sep 21, 2023 · Mnemonic: “sln is the first semisimple Lie algebra you ever learn about, so its Dynkin diagram comes first in the alphabet.” • Weyl group: Sn.
  23. [23]
    [PDF] Introductory Lie Theory Notes - UC Berkeley math
    There exists a unique. Lie group morphism γx : K → G such that ˙γx(0) = x. This map is called the one-parameter subgroup of x. Proof. It is easy to see that ˙γx ...
  24. [24]
    Sophus Lie (1842 - 1899) - Biography - MacTutor
    It was during the winter of 1873-74 that Lie began to develop systematically what became his theory of continuous transformation groups, later called Lie groups ...
  25. [25]
    Essays in the history of Lie groups and algebraic groups, by Armand ...
    Feb 12, 2003 · In the beginning, Sophus Lie hoped to develop a Galois theory for differential equations. Felix Klein's famous “Erlangen Program,” which ...Missing: continuous | Show results with:continuous
  26. [26]
    Wilhelm Killing (1847 - 1923) - Biography - MacTutor
    The main tools in the classification of the semisimple Lie algebras are Cartan subalgebras and the Cartan matrix both first introduced by Killing. He also ...
  27. [27]
    Élie Cartan (1869 - 1951) - Biography - MacTutor
    ... classification of the semisimple algebras over the complex field which Killing had essentially found. ... simple Lie algebras over the complex field. His ...
  28. [28]
  29. [29]
    Hermann Weyl (1885 - 1955) - Biography - MacTutor
    From 1923-38 Weyl evolved the concept of continuous groups using matrix representations. In particular his theory of representations of semisimple groups ...
  30. [30]
    Claude Chevalley (1909 - 1984) - Biography - MacTutor
    Chevalley groups play a central role in the classification of finite simple groups. ... But Chevalley's approach to Clifford algebras was quite new in the 1950s ...
  31. [31]
    Harish-Chandra (1923 - 1983) - Biography - MacTutor
    He has built a fundamental theory of representations of Lie groups and Lie algebras, respectively of harmonic analysis on these groups and their homogeneous ...<|control11|><|separator|>
  32. [32]
    Introduction to Lie Algebras - American Mathematical Society
    Jan 16, 2025 · The treatise [Haw00] by Thomas Hawkins is the definitive and essential work on the history of Lie theory. A great deal of what I say about that ...<|control11|><|separator|>
  33. [33]
    [PDF] Lecture 17 — Cartan Matrices and Dynkin Diagrams
    Nov 4, 2010 · We'll see that to every root system there corresponds a semisimple Lie algebra, so it's important to know all the root systems. Last time we saw ...
  34. [34]
    None
    ### Summary of Differences Between Dynkin and Coxeter Diagrams in Lie Algebras
  35. [35]
    [PDF] Lie algebras - Universidad de Zaragoza
    These notes are intended to provide an introduction to the basic theory of finite dimensional Lie algebras over an algebraically closed field of characteristic ...
  36. [36]
    [PDF] Classification of Real Forms of Semisimple Lie Algebras
    Classification of real forms of semisimple Lie algebras. 41.1. The compact real form. An important example of a real form of simple complex Lie algebra g is ...
  37. [37]
    [PDF] Compact Lie Groups - University of Oregon
    May 5, 2022 · These notes are about Lie groups which are compact, and about the irreducible represen- tations of these groups. In our sketch of the ...
  38. [38]
    [PDF] Classification of Compact Simple Lie Algebras - ETH Zürich
    May 12, 2018 · In this reprt, the compact Lie algebras are classified via the classification of complex simple Lie algebras.
  39. [39]
    [PDF] Lecture Notes on Compact Lie Groups and Their Representations
    ... sp(n,C) is a complex simple Lie algebra and sp(n) is a compact simple Lie algebra. 5.4.19 Example (Adjoint representation of types Bn and Dn) Let g = so(m,C).<|control11|><|separator|>
  40. [40]
    [PDF] math 210c. compact lie groups - Harvard Mathematics Department
    The goal is to cover the structure theory of semisimple compact connected Lie groups, an excellent “test case” for the general theory of semisimple ...<|control11|><|separator|>
  41. [41]
    Introduction to Lie Algebras and Representation Theory - SpringerLink
    In stockDownload chapter PDF · Basic Concepts. James E. Humphreys. Pages 1-14. Semisimple Lie Algebras. James E. Humphreys. Pages 15-41. Root Systems. James E.
  42. [42]
    [PDF] Proof of the Peter-Weyl Theorem - MIT OpenCourseWare
    Exercise 36.8. Show that any compact Lie group admits a structure of a metric space such that the metric is invariant under left and right translations.
  43. [43]
    [PDF] Lie Groups - Berkeley Math
    May 25, 2012 · A typical example of a solvable Lie group is the group of upper triangular matrices with nonzero determinant. (Recall that solvable means the ...
  44. [44]
    [PDF] arXiv:1708.02802v1 [math.CV] 9 Aug 2017
    Aug 9, 2017 · Every complex semisimple Lie group S admits a non-tame discrete subset. Proof. S is Stein as a complex manifold. Due to [6] there exists a dis-.
  45. [45]
    a quick proof of the classification of simple real lie algebras
    Abstract. Élie Cartan's classification of the simple Lie algebras over R is derived quickly from some structure theory over R and the classification over. C.
  46. [46]
    [PDF] Classification of complex and real semisimple Lie Algebras
    We will identify two special real forms of complex semisimple Lie algebras now. The first one will be called split real form. Let g be a complex semisimple ...
  47. [47]
  48. [48]
    [PDF] 37 Simple real Lie algebras - UC Berkeley math
    So we have constructed 3 + 4 + 5 + 3 + 2 (from E8, E7, E6, F4, G2) real forms of exceptional simple Lie groups. There are another 5 exceptional real Lie groups: ...<|control11|><|separator|>
  49. [49]
    [PDF] The Story of sl(2,C) and its Representations or Watch Charlotte ...
    Aug 2, 2012 · We will compute the Lie algebra su(2) of SU(2), complexify su(2) to get sl(2,C), the Lie algebra of SL(2,C), and since we may pass easily be-.
  50. [50]
    Visualizing Lie Subalgebras using Root and Weight Diagrams
    Both diagrams indicate that the dimension of $A_2 = su(3)$ is $8 = 6+2$. A complex semi-simple Lie algebra can almost be identified by its dimension and rank.
  51. [51]
    [PDF] 18.745 F20 Lecture 06: Classical Lie Groups - MIT OpenCourseWare
    An important role in the theory of Lie groups is played by the algebra of quaternions, which is the only noncommu- tative finite dimensional division algebra ...
  52. [52]
    [PDF] Lie-Groups und Representation Theory Exercise Sheet 12
    Jul 9, 2019 · . (c) Show that there is an isomorphism SU(4) → Spin(6). ... (d) Show that this isomorphism restricts to an isomorphism Sp(2) → Spin(5).
  53. [53]
    [PDF] Exceptional Groups and Their Modular Forms - UCSD Math
    The other exceptional groups are labeled 𝐹4, 𝐸6, 𝐸7, 𝐸8. They sit in a chain 𝐺2 ⊆ 𝐹4 ⊆. 𝐸6 ⊆ 𝐸7 ⊆ 𝐸8, and are of dimensions 14, 52, 78, 133, 248 ...<|separator|>
  54. [54]
    [PDF] e8, the most exceptional group - Skip Garibaldi
    The five exceptional simple Lie algebras over the complex number are included one within the other as g2 ⊂ f4 ⊂ e6 ⊂ e7 ⊂ e8. The biggest one, e8, is in many ...
  55. [55]
    Deriving the exceptional Lie groups G2, F4 and E6 - ScienceDirect
    Dec 10, 2023 · The smallest exceptional Lie group is the G 2 group, which has rank 2 and dimension 14. This group emerges as the automorphism group of the ...
  56. [56]
    [PDF] Real Forms of Exceptional Lie Algebras - MIT OpenCourseWare
    It turns out that both for E6 and E8 these give two different real forms, one of which is ... Let g be the complex Lie algebra of type G2, and G the.
  57. [57]
    [PDF] Understanding Lie groups G2 and F4 through the Octonions
    Every simple, compact, connected, and simply connected Lie group is isomorphic to a special unitary group SU(n); a symplectic group Sp(n); a spin group Spin(n),.Missing: A_n B_n C_n D_n
  58. [58]
    [0705.3978] Mapping the geometry of the F4 group - arXiv
    May 28, 2007 · We realize F4 as the automorphisms group of the exceptional Jordan algebra, whose elements are 3 x 3 hermitian matrices with octonionic entries.
  59. [59]
    Models of the Lie algebra F4 - ScienceDirect.com
    We describe the models of the exceptional Lie algebra F 4 which are based on its semisimple subalgebras of rank 4. The underlying fact is that any reductive ...
  60. [60]
    CLASSIFICATION OF PARA-REAL FORMS OF ABSOLUTELY ...
    May 19, 2020 · EVII e7(−25). —. Page 23. Para-Real Forms of Para-Hermitian Symmetric Spaces. 581. We setup the following notations: M(p,q)(K) : the set of all ...
  61. [61]
    [PDF] A survey of six ADE classifications - DPMMS
    We say a simple Lie algebra g is simply laced if the associated root system under Theorem 4.4 is simply laced,. i.e. corresponds to an ADE diagram. Example ...
  62. [62]
    [PDF] Chapter 1 ROOT SYSTEMS AND THEIR CLASSIFICATION - INFN
    So the Cartan matrix is symmetric only if all the simple roots have the same length (in which case the algebra is said to be a simply-laced Lie algebra. Example ...
  63. [63]
    [PDF] The eigenvalues of the quadratic Casimir operator and second ...
    For a simply laced Lie algebra (defined as a simple Lie algebra whose roots are all of equal length), we have g = h. The dual Coxeter number can be related to ...
  64. [64]
    [PDF] on simply laced lie algebras and their minuscule representations
    The Lie algebra E6 may be defined as the algebra of endomorphisms of a 27-dimensional complex vector space MC which annihilate a particular cubic polynomial.Missing: ADE | Show results with:ADE<|control11|><|separator|>
  65. [65]
    [PDF] Basics on Hermitian Symmetric Spaces
    Further we give some examples and deduce from the definition that a symmetric space has the form. G/K for a Lie group G (the automorphism group) and a compact ...
  66. [66]
    Joseph A. Wolf - Scientific Publications - UC Berkeley math
    On the classification of hermitian symmetric spaces. Journal of Mathematics and Mechanics, vol. 13 (1964), pp. 489-496. link; 18. Differentiable fibre spaces ...
  67. [67]
    [PDF] Hermitian symmetric spaces and Kähler rigidity - ETH Zürich
    The stabilizer K := StabG(x0) of a given point x0 ∈ X is a maximal compact subgroup of G and the Lie algebra g of G decomposes as a direct sum g = k⊕p, where k ...