Fact-checked by Grok 2 weeks ago

Commutator

In , the commutator is a that measures the degree to which two elements fail to commute under a given , typically defined as [a, b] = aba^{-1}b^{-1} in groups or [a, b] = ab - ba in associative algebras and related contexts. This concept is central to , where the [G, G] of a group G—generated by all commutators—forms the smallest such that the G/[G, G] is abelian, providing a key tool for studying non-abelian groups and their derived series. Elements a and b are said to commute if [a, b] equals the , highlighting the operation's role in classifying commutative versus non-commutative structures. In the theory of Lie algebras, the commutator serves as the Lie bracket, a bilinear, skew-symmetric operation [x, y] = xy - yx on a that also satisfies the [x, [y, z]] + [y, [z, x]] + [z, [x, y]] = 0, underpinning the local structure of Lie groups and applications in , , and physics. Lie algebras derived from matrix commutators, such as \mathfrak{sl}(n), are particularly influential in symmetry studies and . In , the commutator of two observables represented by Hermitian operators A and B is [A, B] = AB - BA; if this vanishes, the observables commute and can be simultaneously measured with arbitrary precision, whereas non-zero commutators imply fundamental limits via the Heisenberg uncertainty principle, such as [\hat{x}, \hat{p}] = i\hbar for and . This framework extends to broader operator algebras and underpins much of , including and condensed matter systems. Beyond and physics, the term commutator also denotes a mechanical component in : a segmented cylindrical switch in direct-current () motors and generators that reverses the flow of in the armature windings to sustain unidirectional or output voltage. This , essential to brushed DC machines since the early , enables practical electromechanical but introduces challenges like wear and sparking.

Group Theory

Definition

In group theory, the commutator of two elements g, h \in G in a group G is defined as [g, h] = g^{-1} h^{-1} g h. This expression quantifies the extent to which g and h fail to commute, with [g, h] = e (the ) if and only if g and h commute.

Identities and Properties

In , the commutator operation satisfies several basic identities that highlight its . One fundamental identity is that the inverse of a commutator is obtained by swapping the arguments: [g, h]^{-1} = [h, g] for all elements g, h in a group G. Another key identity concerns the interaction with products: [g, hk] = [g, k][g, h]^k, where ^k denotes conjugation by k, i.e., a^k = k^{-1}ak. These identities can be verified by direct into the definition [g, h] = g^{-1}h^{-1}gh and expanding using the group axioms. A significant higher-order relation is the Hall–Witt identity, which extends commutator properties to three elements. If x, y, z \in G satisfy x y z = e, then [[x, y^{-1}], z]^y \cdot [[y, z^{-1}], x]^z \cdot [[z, x^{-1}], y]^x = e. This three-variable identity captures intricate dependencies among commutators and plays a crucial role in analyzing group and extensions. The [G, G], generated by all commutators in G, exhibits important structural properties. By definition, every single commutator [g, h] lies in [G, G]. Furthermore, [G, G] is in G, since conjugating a commutator yields another commutator: g[g', h']g^{-1} = [g g' g^{-1}, g h' g^{-1}] for all g, g', h' \in G. In groups, iterated commutators eventually vanish; specifically, the lower central series G = \gamma_1(G) \triangleright \gamma_2(G) \triangleright \cdots, where \gamma_{i+1}(G) = [\gamma_i(G), G], terminates at the trivial \{e\} after finitely many steps. As an illustrative example, consider s: in the F on a finite or countable generating set, the commutators generate the derived [F, F].

Derived Subgroup

The derived of a group G, denoted [G, G] or G', is the generated by all elements of the form [g, h] = g^{-1} h^{-1} g h for g, h \in G. This is normal in G, and the G / [G, G] is the largest abelian of G, known as the abelianization. The derived series of G is the descending chain of subgroups defined recursively by G^{(0)} = G and G^{(n+1)} = [G^{(n)}, G^{(n)}] for n \geq 0. A group G is solvable if its derived series terminates at the trivial subgroup, meaning there exists some finite k such that G^{(k)} = \{e\}; the smallest such k is called the derived length or solvability length of G. This series captures the extent to which G deviates from being abelian, with each step factoring out commutators to progressively simplify the structure. Nilpotency relates to the derived subgroup through the lower central series, defined by \gamma_1(G) = G and \gamma_{k+1}(G) = [G, \gamma_k(G)] for k \geq 1. A group G is nilpotent if this series reaches the trivial subgroup in finitely many steps, i.e., \gamma_m(G) = \{e\} for some m; the smallest such m is the nilpotency class. Unlike the derived series, which iterates commutators within the previous term, the lower central series incorporates commutators with the full group G, providing a finer measure of "near-commutativity" that implies solvability but not conversely. For abelian groups, the derived subgroup is trivial, as all commutators equal the . In contrast, the A_n for n \geq 5 has derived subgroup [A_n, A_n] = A_n, since A_n is a non-abelian and thus admits no proper nontrivial normal s, forcing the normal derived subgroup to coincide with the whole group. The derived subgroup and its associated series play a key role in the of finite groups, particularly in distinguishing solvable groups—whose composition factors are cyclic of prime order—from nonsolvable ones, aiding the decomposition in the . Historically, these concepts informed the from the early , which asks whether finitely generated groups of bounded exponent are finite; negative solutions in the mid-20th century relied on constructing infinite groups via commutator relations in free Burnside groups.

Ring Theory

Definition

In ring theory, the commutator of two elements a, b \in R in an associative ring R (not necessarily commutative or unital) is defined as [a, b] = ab - ba. This expression quantifies the extent to which multiplication in R fails to be commutative, analogous to the role of the group commutator in measuring deviations from multiplicativity in groups. The commutator operation is skew-symmetric, satisfying [a, b] = -[b, a] for all a, b \in R, which follows directly from the definition. It is also bilinear over the integers, meaning [\lambda a + \mu c, b] = \lambda [a, b] + \mu [c, b] and [a, \lambda b + \mu d] = \lambda [a, b] + \mu [a, d] for \lambda, \mu \in \mathbb{Z}, due to the bilinearity of multiplication in associative rings. A concrete example arises in the ring M_n(F) of n \times n matrices over a field F, where the standard matrix units E_{ij} (with 1 in the (i,j)-position and zeros elsewhere) satisfy [E_{ij}, E_{kl}] = \delta_{jk} E_{il} - \delta_{li} E_{kj}, with \delta denoting the Kronecker delta. This computation illustrates how commutators generate the special linear Lie algebra \mathfrak{sl}_n(F) as the trace-zero matrices.

Basic Identities

In , the commutator operation [a, b] = ab - ba exhibits linearity with respect to in both arguments. Specifically, for all a, b, c \in R, [a + b, c] = [a, c] + [b, c], \quad [a, b + c] = [a, b] + [a, c]. These identities follow directly from the bilinearity of over in any associative R. The commutator also obeys a resembling the Leibniz rule for derivations: [ab, c] = a[b, c] + [a, c]b, \quad [a, bc] = [a, b]c + b[a, c] for all a, b, c \in R. These can be verified by expanding the definitions using the associativity and distributivity axioms of rings. The centralizer of R, consisting of elements that commute with every element of R, is the center Z(R) = \{ z \in R \mid [z, r] = 0 \ \forall r \in R \}. This is always a two-sided of R. The map ad_a : r \mapsto [a, r] then defines an inner on R for each a \in R. As an example, consider the polynomial ring k over a field k. Since k is commutative, all commutators vanish, so [f, g] = 0 for any polynomials f, g \in k, regardless of degrees.

Advanced Identities

In matrix algebras over a field, the trace of a commutator vanishes, i.e., for any matrices A and B, \operatorname{tr}([A, B]) = 0. This follows from the cyclic property of the trace, \operatorname{tr}(AB) = \operatorname{tr}(BA), yielding \operatorname{tr}([A, B]) = \operatorname{tr}(AB - BA) = \operatorname{tr}(AB) - \operatorname{tr}(BA) = 0. A key identity for powers in associative rings is the generalized Leibniz rule for commutators: for elements a, b \in R and nonnegative integer n, [a, b^n] = \sum_{k=0}^{n-1} b^k [a, b] b^{n-1-k}. This holds by on n, with the base case n=1 trivial, and the inductive step using the [a, bc] = [a, b]c + b[a, c], which commutators satisfy in any associative ring. In rings admitting exponentials, such as those over fields of characteristic zero or matrix rings, the exponential of a commutator relates to conjugations via the . Specifically, for elements a, b where higher nested commutators are negligible (e.g., when \|[a, b]\| is small), e^{[a, b]} \approx e^a e^b e^{-a} e^{-b}, with the arising from the e^a e^b e^{-a} e^{-b} = e^{[a, b] + \frac{1}{2}[a, [a, b]] + \cdots}. These identities underpin the definition of the commutator [R, R] in a ring R, the two-sided generated by all commutators \{ [a, b] \mid a, b \in R \}. This captures the "noncommutativity" of R, and the R / [R, R] is the largest commutative .

Lie Algebras

Definition as Lie Bracket

In a \mathcal{L} over a K, the commutator is defined as the Lie bracket [x, y] for elements x, y \in \mathcal{L}, satisfying the antisymmetry condition [x, y] = -[y, x]. This bracket operation turns \mathcal{L} into a non-associative structure, where the commutator captures the failure of elements to commute, distinct from the multiplicative structure in associative algebras. The Lie bracket often arises from associative algebras by imposing the commutator [x, y] = xy - yx on the underlying vector space, where xy denotes the original associative product. This construction endows the vector space with a structure, preserving antisymmetry as a direct consequence of the . Such from associative origins, like algebras, illustrate how the commutator serves as a fundamental operation in the study of symmetries. The Lie bracket is bilinear over K, meaning [a x + b y, z] = a [x, z] + b [y, z] and [z, a x + b y] = a [z, x] + b [z, y] for all scalars a, b \in K and elements x, y, z \in \mathcal{L}. This bilinearity ensures the bracket behaves linearly with respect to operations, facilitating the algebraic manipulations central to . A canonical example is the general linear \mathfrak{gl}(n, K), consisting of n \times n matrices over K with the bracket [A, B] = AB - BA. Another is the Heisenberg algebra, a three-dimensional over K with basis \{x, y, z\} where [x, y] = z and all other brackets vanish, demonstrating a nilpotent structure arising from the commutator.

Jacobi Identity

The Jacobi identity serves as a defining associator for Lie algebras, encapsulating the non-associative nature of the Lie bracket while ensuring structural integrity. For all elements x, y, z in a Lie algebra \mathcal{L}, it states [x, [y, z]] + [y, [z, x]] + [z, [x, y]] = 0. This axiom, alongside bilinearity and antisymmetry, distinguishes Lie algebras from more general non-associative algebras and is essential for modeling symmetries in Lie groups. When constructing algebras from s, the emerges directly from the underlying associativity. In an with product ab, the is the commutator [a, b] = ab - ba. Expanding the left-associated term [[x, y], z] = (xy - yx)z - z(xy - yx) yields xyz - yxz - zxy + zyx, and for the other terms shows that their sum vanishes due to the associative law (ab)c = a(bc), thereby verifying the for the induced structure. Among its key consequences, the Jacobi identity implies that the adjoint representation \mathrm{ad}_x(y) = [x, y] defines a of \mathcal{L}, obeying the Leibniz rule \mathrm{ad}_x[y, z] = [\mathrm{ad}_x y, z] + [y, \mathrm{ad}_x z] for all x, y, z \in \mathcal{L}. This derivation property facilitates the study of ideals, subalgebras, and representations, while also ensuring alternativity in iterated brackets, such as [x, [x, y]] = 0, which follows from antisymmetry but is reinforced by the cyclic consistency of the identity. A concrete illustration occurs in the special linear Lie algebra \mathfrak{sl}(2, K) over a K of characteristic zero, with H = \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix}, X = \begin{pmatrix} 0 & 1 \\ 0 & 0 \end{pmatrix}, Y = \begin{pmatrix} 0 & 0 \\ 1 & 0 \end{pmatrix}. The commutation relations are [H, X] = 2X, [H, Y] = -2Y, [X, Y] = H. To verify the for these basis elements, compute [H, [X, Y]] + [X, [Y, H]] + [Y, [H, X]]: the first term is [H, H] = 0; the second is [X, 2Y] = 2[X, Y] = 2H; the third is [Y, 2X] = 2[Y, X] = -2H. The sum $0 + 2H - 2H = 0 confirms the identity holds, and similar checks for other combinations follow by bilinearity.

Adjoint Representation

In Lie algebras, the adjoint map associated to an element x \in \mathfrak{L} is the linear \mathrm{ad}_x: \mathfrak{L} \to \mathfrak{L} defined by \mathrm{ad}_x(y) = [x, y] for all y \in \mathfrak{L}, where [\cdot, \cdot] denotes the bracket. This map is a of the , meaning it satisfies \mathrm{ad}_x([y, z]) = [\mathrm{ad}_x(y), z] + [y, \mathrm{ad}_x(z)], which follows directly from the . The adjoint representation of the Lie algebra \mathfrak{L} is the homomorphism \rho: \mathfrak{L} \to \mathfrak{gl}(\mathfrak{L}) given by \rho(x) = \mathrm{ad}_x, where \mathfrak{gl}(\mathfrak{L}) is the Lie algebra of linear endomorphisms of \mathfrak{L} equipped with the commutator bracket. The kernel of \rho is the center Z(\mathfrak{L}) = \{ x \in \mathfrak{L} \mid [x, y] = 0 \ \forall y \in \mathfrak{L} \}. A key property is that \mathrm{ad}_{[x, y]} = [\mathrm{ad}_x, \mathrm{ad}_y], where the bracket on the right is the commutator in \mathfrak{gl}(\mathfrak{L}); this confirms that \rho (or \mathrm{ad}) is a Lie algebra homomorphism from \mathfrak{L} to \mathfrak{gl}(\mathfrak{L}). For a concrete example, consider the \mathfrak{so}(3), which consists of $3 \times 3 skew-symmetric real matrices and can be identified with \mathbb{R}^3 under the as the Lie bracket, so that [x, y] = x \times y. In this identification, the acts as \mathrm{ad}_x(y) = x \times y, corresponding to the natural action of rotations on vectors. The Killing form on \mathfrak{so}(3), defined by B(x, y) = \mathrm{tr}(\mathrm{ad}_x \mathrm{ad}_y), is non-degenerate and proportional to the negative of the standard inner product on \mathbb{R}^3.

Advanced Applications

Graded Algebras

In a R = \bigoplus_{i \in I} R_i, where I is an , the commutator of homogeneous elements a \in R_i and b \in R_j is defined as [a, b] = ab - ba. Since the multiplication in R respects the grading, with ab \in R_{i+j} and ba \in R_{j+i} = R_{i+j}, the commutator [a, b] also lies in R_{i+j}. This graded commutator preserves the decomposition and forms the basis for structures in graded settings. For \mathbb{Z}/2\mathbb{Z}-graded algebras, known as superalgebras, the structure is A = A_0 \oplus A_1, where A_0 consists of even elements and A_1 of odd elements, with multiplication satisfying A_i A_j \subseteq A_{i+j \mod 2}. The supercommutator, which generalizes the usual commutator to account for , is defined for homogeneous elements as [a, b]_s = ab - (-1)^{|a||b|} ba, where |a| denotes the (0 or 1) of a; this extends bilinearly to all elements. In Lie superalgebras, the supercommutator serves as the Lie bracket, ensuring antisymmetry up to sign: [b, a]_s = - (-1)^{|a||b|} [a, b]_s. A key property of the supercommutator in Lie superalgebras is the super Jacobi identity, which adapts the classical to incorporate grading: for homogeneous elements x, y, z, (-1)^{|x||y|} [[x, y]_s, z]_s + (-1)^{|y||z|} [[y, z]_s, x]_s + (-1)^{|z||x|} [[z, x]_s, y]_s = 0. This identity ensures the bracket defines a consistent and holds in any superalgebra where the supercommutator is used. Applications of supercommutators appear prominently in , where supercharges Q_\alpha satisfy graded commutation relations like \{ Q_\alpha, \bar{Q}_\beta \} = 2 \sigma^\mu_{\alpha \beta} P_\mu, closing the algebra on spacetime translations and the , thus unifying bosonic and fermionic symmetries. An illustrative example is the \mathrm{Cl}(V, Q), which is \mathbb{Z}/2\mathbb{Z}-graded with even and odd parts, where the algebra is generated by vectors e_i satisfying the anticommutator relation \{ e_i, e_j \} = 2 g_{ij} \mathbf{1} (with g the metric) for the defining Q. The supercommutator of two odd generators is [e_i, e_j]_s = e_i e_j + e_j e_i = \{ e_i, e_j \} = 2 g_{ij} \mathbf{1}, which is central (even). However, the ordinary commutator [e_i, e_j] = e_i e_j - e_j e_i = 2 (e_i e_j - g_{ij} \mathbf{1}) lies in the even and generates the \mathfrak{so}(V), linking to rotations and representations in the even .

Derivations and Adjoint Derivation

In ring theory, a derivation on a ring R is a linear map \delta: R \to R that satisfies the Leibniz rule \delta(ab) = \delta(a)b + a\delta(b) for all a, b \in R, along with additivity \delta(a + b) = \delta(a) + \delta(b). This structure generalizes the familiar product rule from calculus to abstract algebraic settings. An adjoint derivation, also known as an inner derivation, arises from the commutator operation within the itself; specifically, for a fixed a \in [R](/page/R), the map \mathrm{ad}_a: [R](/page/R) \to [R](/page/R) defined by \mathrm{ad}_a(b) = [a, b] = ab - ba is a . The set \mathrm{Inn}(R) of all such inner derivations forms a subring of the \mathrm{Der}(R) of all derivations on R. Derivations that cannot be expressed as inner derivations for any a \in R are termed outer derivations. The quotient space \mathrm{Der}(R)/\mathrm{Inn}(R) provides a classification of outer derivations up to equivalence by inner ones, capturing the "essential" non-inner behavior of derivations on R. A notable example occurs in the Weyl algebra, the generated by a and a operator; here, the acts as an outer on the underlying commutative , since inner derivations vanish in the commutative case. This construction via Ore extension using an outer realizes the ring of polynomial differential operators. In the context of Lie algebras, the adjoint derivation \mathrm{ad}_a serves as a special case restricted to the Lie bracket.

References

  1. [1]
    Commutator -- from Wolfram MathWorld
    The commutator of two group elements A and B is ABA^(-1)B^(-1) , and two elements A and B are said to commute when their commutator is the identity element.
  2. [2]
    [PDF] Definition 1.10. If S is a subset of a group G then the subgroup ...
    Definition 1.11. The commutator subgroup (also called the derived subgroup) G# = [G, G] is the subgroup of G generated by all commuta- tors [a, b].<|separator|>
  3. [3]
    [PDF] The Commutator Subgroup
    If G is Abelian, then we have C = {e}, so in one sense the commutator subgroup may be used as one measure of how far a group is from being Abelian. Specifically ...
  4. [4]
    [PDF] Lecture 1 - Basic Definitions and Examples of Lie Algebras
    Sep 6, 2012 · A Lie algebra l is a vector space V over a base field F, along with an operation [·, ·] : V ×V →. V called the bracket or commutator that ...
  5. [5]
    [PDF] Representations of Matrix Lie Algebras - UChicago Math
    The standard Lie bracket for matrix Lie algebras, called the commutator bracket, is defined for A, B ∈ g as [A, B] = AB − BA. 2. Page 3. 2 The Lie Algebra of a ...
  6. [6]
    4.5 The Commutator - FAMU-FSU College of Engineering
    4.5 The Commutator ... is a measure of the uncertainty of a property of a quantum system. The larger the standard deviation, the farther typical measurements ...
  7. [7]
    [PDF] Lecture 7 - MIT OpenCourseWare
    The Commutator operator is defined as: ⌈Aˆ, B ... In fact, every symmetry in a physical system implies an associated conserved quantity. In quantum mechanics the ...<|control11|><|separator|>
  8. [8]
    [PDF] DC (Commutator) and Permanent Magnet Machines
    The starting motor on all automobiles is a series-connected commutator machine. Many of the other electric motors in automobiles, from the little motors that ...
  9. [9]
    [PDF] Basics on electric motors
    Apr 29, 2010 · The split ring commutators joins both ends of the loop to fixed wires, and every half-rotation, the contacts are switched. In a generator, the ...
  10. [10]
    ON THE PROJECTIVE COMMUTANT OF ABELIAN GROUPS
    Recall that if R is a ring and a, b ∈ R then the element. [a, b] = ab−ba is called the commutator of a and b. ... Then the ring E(B) is commutative, and since. B ...
  11. [11]
    Commutators in associative rings | Mathematical Proceedings of the ...
    Oct 24, 2008 · Let R be an arbitrary associative ring, and X a set of generators of R. The elements of X generate a Lie ring, [X], say, with respect to the ...Missing: bilinearity | Show results with:bilinearity
  12. [12]
    [PDF] Matrix Calculus and Kronecker Product - NET
    (Eij)(Ekl) = δjk(Eil) it follows that the commutator is given by. [(Eij), (Ekl)] = δjk(Eil) − δli(Ekj) . Thus the coefficients are all ±1 or 0. The ...
  13. [13]
    Finite Groups - Daniel Gorenstein - Google Books
    This volume provides a relatively concise and readable access to the key ideas and theorems underlying the study of finite simple groups and their important ...Missing: URL | Show results with:URL
  14. [14]
    [PDF] The commutator subgroups of free groups and surface groups
    1 Introduction. Let Fn be the free group on {x1,...,xn}. The commutator subgroup [Fn,Fn] is an infinite-rank free group.
  15. [15]
    [PDF] SOLVABLE AND NILPOTENT GROUPS
    G0 = G, Gn+1 = [G, Gn]. This second series is called the lower central series of G. Clearly, both the G(n) and the Gn are fully invariant subgroups of G. ...
  16. [16]
    [PDF] The Alternating Group
    An is called the alternating group. An important feature of the alternating group is that, unless n = 4, it is a simple group. A group G is said to be simple ...
  17. [17]
    [PDF] On the Classification of Finite Simple Groups - MIT Mathematics
    May 22, 2022 · This paper examines the properties of finite simple groups, which arise from the decomposition of groups into normal subgroups and a quotient.
  18. [18]
    [PDF] The Burnside Problem - Marcel Goh
    Apr 15, 2020 · Introduction. ONE OF THE oldest and most studied problems in group theory began life in a 1902 paper by William. Burnside [3].
  19. [19]
    A First Course in Noncommutative Rings - Google Books
    A First Course in Noncommutative Rings. Front Cover. Tsit-Yuen Lam. Springer Science & Business Media, Jun 21, 2001 - Mathematics - 385 pages. A First Course in ...
  20. [20]
    [PDF] On Trace Zero Matrices and Commutators - arXiv
    Nov 8, 2021 · This will be discussed further in Section 6 along with Theorem 6.11 which characterises rings where every 2 × 2 trace 0 matrix is a commutator.
  21. [21]
    [PDF] Ode to commutator operators - arXiv
    Nov 23, 2007 · Commutators automatically satisfy the Leibniz rule for products, [a, b c] = abc - bca = abc - bac + bac - bca (2.2) = [a, b] c + b [a, c]. Thus ...<|separator|>
  22. [22]
    [PDF] Notes on Baker-Campbell-Hausdorff (BCH) Formulae - Duke Physics
    An important physics application of this form of the BCH formulae is in the theory of coherent states. In general, there is no closed form expression for the ...
  23. [23]
    [PDF] Commutator rings∗ - UCCS
    Jul 6, 2006 · Commutator rings∗ Zachary Mesyan. July 6, 2006. Abstract. A ring is called a commutator ring if every element is a sum of additive commutators.
  24. [24]
    [PDF] Lie algebras and their root systems
    A Lie algebra L is a vector space with a skew-symmetric bilinear map, called the. Lie bracket or commutator and written as [·, ·]: L × L → L, which satisfies ...
  25. [25]
    [PDF] math 223a notes 2011 lie algebras - Brandeis
    Sep 6, 2011 · Then the commutator [xy] is defined by [xy] = xy − yx. This is easily seen to be a bracket and is also called the Lie bracket of the associative ...<|separator|>
  26. [26]
    [PDF] Lie Algebras: Definition and Basic Properties
    The Lie bracket [x, y] defined above is called the commutator product of x and y. The associative algebra Fn of n×n matrices can therefore be given a Lie ...
  27. [27]
    [PDF] Lie algebras 1 1.1. Definition 1 1.2. Motivation
    The Jacobi identity holds for any associative algebra (see below) and is left as an exercise. D. Example 1.4. Let R = C, S = C∞(R), the ring of complex ...
  28. [28]
    [PDF] Introduction to Lie Algebras and Representation Theory
    By the Jacobi identity,. N(K) is a subalgebra of L; it may be described verbally as the largest sub- algebra of L which includes K as an ideal (in case K is ...
  29. [29]
    [PDF] 1. The Lie algebra sl2 and its finite dimensional representations
    The proofs follow in the later subsections below. (a) (“complete reducibility”) every such representation is a direct sum of irre- ducible sub representations.
  30. [30]
    The Adjoint Representation - BOOKS
    Section 5.1 The Adjoint Representation ... Any Lie algebra g acts on itself via commutators. This action is linear, so we can represent it using matrices. ... in ...
  31. [31]
    [PDF] Contents - Physics Courses
    For example, the vector space R3 endowed with the vector cross product as the Lie bracket is a Lie algebra. The corresponding Lie group is SO(3).
  32. [32]
    [PDF] Lie Groups and Algebras 1 Intro 2 The Adjoint Representation and ...
    Jan 6, 2016 · representing A and B in the adjoint representation: (A, B) := Tr(DA ... A Lie algebra is determined by its structure constants but this ...
  33. [33]
    Supercommutator algebras of right (Hom-)alternative superalgebras
    Oct 7, 2017 · The supercommutator algebra of a right alternative superalgebra is a Bol superalgebra. Hom-Bol superalgebras are defined and it is shown that they are closed ...
  34. [34]
    [PDF] Jacobi - type identities in algebras and superalgebras - arXiv
    Jun 6, 2014 · For any algebra, we prove that if the fundamental identity is satisfied, then the multiplication operation is associative. We generalize the ...
  35. [35]
    [PDF] Supersymmetry Basics Contents - Zhong-Zhi Xianyu
    This is the first piece of a series of notes on supersymmetry. We intend to present the very basics of supersymmetry that is needed to explore more advanced ...
  36. [36]
    [PDF] Clifford algebras
    The Clifford algebra also carries an ordinary trace, vanishing on ordinary commutators. Proposition 2.11. The formula tr: Cl(V ;B) → K, x 7→ σ(x)[0].
  37. [37]
    (PDF) ON DERIVATIONS IN RINGS AND THEIR APPLICATIONS
    derivations of Rof the form rδ:x→rδ(x), r∈R, is a Lie subring of the Lie ring D(R) of derivations of R. In [135], Jordan & Jordan studied the structure of D(R) ...
  38. [38]
    [PDF] Enveloping Algebras of Derivations of Commutative ... - eScholarship
    Sep 9, 2025 · An inner derivation of A is a derivation of the form ada : A → A defined by ada(b) = ab − ba, where a, b ∈ A. The set of all inner ...
  39. [39]
    [PDF] Derivations and Integrations on Rings
    Feb 9, 2019 · For distinguishing between the two cases, a derivation of R which is not inner is called an outer derivation. 2.2 Definition: Let R be a ring ...
  40. [40]
    [PDF] A flavour of noncommutative algebra (part 1) - Vipul Naik
    Thus, the differentiation in either variable in a polynomial ring is an outer derivation. ... The Weyl algebra. In the next subsection, we shall see “concretely” ...
  41. [41]
    [PDF] Derivations and Iterated Skew Polynomial Rings - arXiv
    first Weyl algebra over T and is denoted by A1(T). It becomes evident that ... an outer derivation of the polynomial ring R=T[x1] over a ring T ...