Fact-checked by Grok 2 weeks ago

*-algebra

A *-algebra (or star-algebra) is an associative algebra over the complex numbers \mathbb{C} equipped with an involution operation ^*, which is a conjugate-linear map satisfying (a + b)^* = a^* + b^*, (\lambda a)^* = \bar{\lambda} a^* for \lambda \in \mathbb{C}, (ab)^* = b^* a^*, and (a^*)^* = a for all a, b in the algebra. This structure introduces a notion of "adjoint" analogous to the Hermitian adjoint in linear algebra, enabling the study of self-adjoint elements (a^* = a) and more general symmetry properties. *-Algebras form the foundational framework for operator algebras in functional analysis, where they model systems with both multiplicative and adjoint structures, such as bounded linear operators on Hilbert spaces equipped with the adjoint operation. Notable examples include the algebra of complex numbers with complex conjugation as the involution, the algebra of n \times n complex matrices with conjugate transpose, and the algebra of continuous functions on a compact space with pointwise conjugation. In unital *-algebras, the unit element satisfies $1^* = 1, and for invertible elements, the inverse of the adjoint equals the adjoint of the inverse: (a^{-1})^* = (a^*)^{-1}. The theory of -algebras extends to normed settings, leading to C-algebras, which are complete Banach -algebras satisfying the C-identity \|a^* a\| = \|a\|^2, crucial for spectral theory and representation theorems. These structures underpin applications in quantum mechanics, where self-adjoint elements represent observables, and in non-commutative geometry, facilitating the algebraic description of spaces without classical topology. Further developments include von Neumann algebras, which are weak closures of *-algebras of operators, essential for understanding infinite-dimensional phenomena in physics and probability.

Definitions

*-rings

A *-ring is a ring R equipped with an involution *: R \to R, which is an anti-automorphism of R satisfying (x + y)^* = x^* + y^*, (xy)^* = y^* x^*, and (x^*)^* = x for all x, y \in R. This structure generalizes the notion of conjugation in familiar examples like the quaternions, where the involution reverses the non-commutative multiplication while preserving addition and being its own inverse. If the *-ring R is unital with multiplicative identity $1, then the involution fixes the identity, i.e., $1^* = 1. This follows from the anti-automorphism property applied to the identity relations $1x = x = x1 for all x \in R, yielding x^* = 1^* x^* = x^* 1^* and thus $1^* = 1. In a *-ring R, the self-adjoint elements are those x \in R satisfying x^* = x, often denoted S(R) = \{x \in R \mid x^* = x\}, while the skew-adjoint elements satisfy x^* = -x, denoted K(R) = \{x \in R \mid x^* = -x\}. Assuming the ring has no 2-torsion (i.e., characteristic not 2), these sets partition R additively via the direct sum decomposition R = S(R) \oplus K(R), where every element decomposes uniquely as x = s + k with s \in S(R), k \in K(R), given by s = \frac{x + x^*}{2} and k = \frac{x - x^*}{2}. The set S(R) of self-adjoint elements forms an additive subgroup of R, as (x + y)^* = x^* + y^* = x + y for x, y \in S(R), and it is closed under the Jordan multiplication defined by x \circ y = \frac{1}{2}(xy + yx), since (x \circ y)^* = \frac{1}{2}(y^* x^* + x^* y^*) = \frac{1}{2}(yx + xy) = x \circ y. Thus, S(R) is a Jordan subring of R under addition and this symmetrized multiplication, capturing the commutative aspects induced by the involution. This structure underlies many applications of *-rings to symmetric forms and representations. *-Rings provide the ring-theoretic foundation for *-algebras, where compatibility with scalar multiplication is additionally required.

*-algebras

A *-algebra is defined as a *-ring A that is also an algebra over a commutative unital *-ring R, equipped with an involution * on A satisfying (r x)^* = r^* x^* for all r \in R and x \in A. This compatibility condition ensures that the scalar multiplication respects the involutive structure of the base ring R. Typically, R is taken to be the complex numbers \mathbb{C} equipped with complex conjugation as its involution, though the definition extends to other *-rings such as the real numbers \mathbb{R} with the trivial involution. When the base ring is \mathbb{C}, the involution on the *-algebra A is conjugate-linear, meaning (\lambda x)^* = \bar{\lambda} x^* for \lambda \in \mathbb{C} and x \in A. In this setting, A is a complex vector space with an associative bilinear multiplication and the anti-automorphism property of the involution inherited from the *-ring structure. Over \mathbb{R}, where the involution on the base is the identity, the map x \mapsto x^* is linear with respect to real scalars. A *-homomorphism between two *-algebras A and B over the same base ring is a ring homomorphism f: A \to B that preserves the involution, satisfying f(x^*) = f(x)^* for all x \in A. Such maps automatically respect the module structure over R due to the algebraic compatibility. In certain contexts, such as when the *-algebra is a division algebra, the involution is unique; for example, quaternion division algebras admit a unique canonical involution of the symplectic type.

Involution properties

The involution * on a *-algebra A is an anti-automorphism, meaning it is additive, (x + y)^* = x^* + y^* for all x, y \in A, reverses multiplication, (xy)^* = y^* x^* for all x, y \in A, and is involutive, (x^*)^* = x for all x \in A. When A is a complex algebra, the involution is conjugate-linear over the base field \mathbb{C}, satisfying (\lambda x + \mu y)^* = \bar{\lambda} x^* + \bar{\mu} y^* for all \lambda, \mu \in \mathbb{C} and x, y \in A. A key derived property is that x^* x is self-adjoint, (x^* x)^* = x^* x, for all x \in A. In contexts where A admits a partial order compatible with the involution, such as C*-algebras, x^* x is positive, meaning it belongs to the positive cone. Furthermore, in finite-dimensional cases, the trace satisfies \operatorname{tr}(x^* x) \geq 0 for all x \in A. The involution induces a direct sum decomposition A = H(A) \oplus i H(A), where H(A) = \{ x \in A \mid x^* = x \} is the real subspace of self-adjoint elements and i H(A) = \{ i x \mid x \in H(A) \} consists of the skew-adjoint elements satisfying x^* = -x. The self-adjoint part H(A) forms a Jordan algebra under the symmetrized product x \circ y = \frac{1}{2} (xy + yx), while the skew-adjoint part i H(A) underlies a Lie algebra structure via the commutator [x, y] = xy - yx.

Motivations and Notation

Philosophical background

The *-operation in algebras extends the concept of complex conjugation from the field of complex numbers to more abstract algebraic structures, providing a mechanism to identify and preserve "real" or self-adjoint elements analogous to real numbers. In the complex numbers, conjugation satisfies \overline{z w} = \bar{w} \bar{z} and \bar{\bar{z}} = z, ensuring that self-conjugate elements (those with z = \bar{z}) form the real line, which underpins notions of positivity and reality in analysis. Similarly, the *-involution on an algebra is designed as an anti-automorphism to mirror this reversal in multiplication, allowing self-adjoint elements (a = a^*) to capture the "real structure" essential for spectral properties and positivity conditions in broader settings. This structure draws direct inspiration from linear algebra on inner product spaces, where the Hermitian adjoint operation T^* on linear operators satisfies \langle T x, y \rangle = \langle x, T^* y \rangle, ensuring that the inner product \langle x, y \rangle = \overline{\langle y, x \rangle} preserves a positive-definite form. The *-operation generalizes this adjoint, imposing involution properties such as (a b)^* = b^* a^* and (a^*)^* = a to maintain compatibility with sesquilinear forms, where self-adjoint operators yield real eigenvalues and positive operators correspond to those expressible as a = b^* b. This analogy facilitates the extension of Hermitian theory to abstract algebras, enabling the study of spectra and representations without reliance on concrete Hilbert spaces. In quantum mechanics, the *-involution plays a foundational role by associating self-adjoint elements with physical observables, whose spectra represent possible measurement outcomes, as formalized in the operator algebra approach to quantum theory. John von Neumann's development of rings of operators on Hilbert spaces highlighted how self-adjoint operators ensure real-valued expectations, aligning mathematical structure with empirical reality in quantum systems. This motivation drove the abstraction to *-algebras, allowing non-commutative geometries to model quantum phenomena beyond finite dimensions. The choice of an anti-automorphism for the *-operation, rather than a standard automorphism, is deliberate to accurately model the adjoint's reversal of multiplication order, distinguishing it from symmetry-preserving maps like automorphisms that would instead satisfy (a b)^* = a^* b^* and fail to capture the sesquilinear duality inherent in inner products and quantum observables.

Standard notations

In *-algebras, the involution operation is conventionally denoted by the superscript star, mapping an element x to x^*, a notation prevalent in operator algebra literature. This star symbol emphasizes the adjoint or conjugate-like nature of the map, aligning with its role in preserving algebraic structure while introducing anti-automorphism properties. In contexts involving bounded linear operators on Hilbert spaces, the dagger symbol \dagger is frequently used instead, denoting the adjoint as x^\dagger; this convention is common in functional analysis and quantum theory texts. For involutions mimicking complex conjugation, particularly in algebras over the reals or in certain ring-theoretic settings, the overline notation \overline{x} is employed. Notational variations arise based on the underlying field or context: the star is standard for complex *-algebras, reflecting the complex conjugate influence, while the bar appears more often in general ring theory or quaternion algebras with standard conjugation. Occasionally, the hash symbol # is used for the involution in specialized algebraic structures. *-Homomorphisms between *-algebras are algebra homomorphisms that preserve the involution, satisfying f(x^*) = f(x)^* for all x, ensuring compatibility with the *-structure. In standard definitions, this preservation is mandatory, though skew variants may involve anti-preservation, where f(x^*) = \overline{f(x)}^* or similar adjustments to account for twisted multiplications. Typographically, in mathematical typesetting with LaTeX, the star is rendered as superscript via x^* or x^{*}, and the dagger as x^\dagger, facilitating clear distinction in printed and digital formats. These conventions enhance readability in dense algebraic expressions.

Examples

Finite-dimensional cases

The field of complex numbers \mathbb{C}, viewed as a 1-dimensional algebra over itself, forms a *-algebra under the involution given by complex conjugation, where for z = a + bi with a, b \in \mathbb{R} and i^2 = -1, the conjugate is z^* = a - bi. This involution satisfies the required properties: it is antilinear ((cz + w)^* = \bar{c} z^* + w^* for c \in \mathbb{C}), an anti-automorphism ((zw)^* = w^* z^*), and of order two ((z^*)^* = z). As a commutative unital *-algebra, \mathbb{C} serves as the prototypical finite-dimensional example, where the self-adjoint elements (fixed by *) are the real numbers \mathbb{R}. A fundamental noncommutative example is the algebra of n \times n complex matrices M_n(\mathbb{C}), which is a finite-dimensional *-algebra over \mathbb{C} equipped with the Hermitian adjoint as the involution: for a matrix A = (a_{ij}), the adjoint is A^* = (\bar{a}_{ji}), the transpose with entrywise complex conjugation. This operation is antilinear, an anti-automorphism ((AB)^* = B^* A^*), and involutive ((A^*)^* = A), making M_n(\mathbb{C}) unital with identity the standard identity matrix. The self-adjoint elements are precisely the Hermitian matrices (A^* = A), which play a key role in spectral theory within this structure. For n=1, this reduces to \mathbb{C} itself. Over the real numbers \mathbb{R}, the quaternion algebra \mathbb{H} provides a 4-dimensional noncommutative division *-algebra with basis \{1, i, j, k\} satisfying i^2 = j^2 = k^2 = -1, ij = k = -ji, jk = i = -kj, and ki = j = -ik. The standard involution is the quaternion conjugate: for q = a + bi + cj + dk with a, b, c, d \in \mathbb{R}, q^* = a - bi - cj - dk, which negates the pure imaginary part. This map is \mathbb{R}-linear (since the base field is real), an anti-automorphism ((pq)^* = q^* p^*), and involutive ((q^*)^* = q), with self-adjoint elements being the real quaternions (purely real scalars). Unlike \mathbb{C}, \mathbb{H} is noncommutative, highlighting how *-structures can extend to higher-dimensional associative algebras without zero divisors. Finite direct sums offer a way to construct new *-algebras from existing ones; for instance, given *-algebras A and B over the same field, their direct sum A \oplus B = \{(a, b) \mid a \in A, b \in B\} inherits componentwise addition and multiplication, with involution (a, b)^* = (a^*, b^*). This preserves the *-algebra axioms: the involution remains antilinear (or linear if over \mathbb{R}), an anti-automorphism, and involutive on the product structure. The construction extends to finite direct sums \bigoplus_{k=1}^m A_k, yielding an m-component *-algebra whose self-adjoint elements are tuples of self-adjoints from each factor, illustrating how *-algebras can be decomposed into simpler finite-dimensional building blocks. For example, M_n(\mathbb{C}) \oplus M_m(\mathbb{C}) combines matrix algebras in this manner.

Infinite-dimensional cases

In infinite-dimensional *-algebras, the structure extends beyond finite bases, often arising in contexts like operator theory and functional analysis, where the involution interacts with infinite-dimensional vector spaces to model physical systems or geometric objects. These algebras provide foundational examples that connect algebraic properties to analytic phenomena without imposing completeness conditions. A prominent non-commutative example is the algebra B(H) of all bounded linear operators on a complex Hilbert space H, equipped with composition as multiplication and the adjoint operation as the involution, defined by \langle T x, y \rangle = \langle x, T^* y \rangle for all x, y \in H and T \in B(H). This structure is infinite-dimensional whenever H is, capturing transformations on infinite-dimensional spaces like \ell^2(\mathbb{N}). In the commutative case, the algebra C(X, \mathbb{C}) of continuous complex-valued functions on a compact Hausdorff space X, with pointwise multiplication and complex conjugation as the involution (f(x))^* = \overline{f(x)}, exemplifies an infinite-dimensional *-algebra when X is infinite, such as the unit circle. This construction highlights how topological spaces encode algebraic involutions through function spaces. Polynomial algebras over \mathbb{C}, such as \mathbb{C} with the involution extending complex conjugation on coefficients (i.e., ( \sum a_k t^k )^* = \sum \overline{a_k} t^k), form another infinite-dimensional example, where the degree spans infinitely many basis elements. More generally, involutive Hopf *-algebras like the group algebra \mathbb{C}[G] of a discrete group G, with basis elements \delta_g satisfying (\delta_g)^* = \delta_{g^{-1}} and convolution multiplication, arise in representation theory and extend to infinite groups like \mathbb{Z}. Tensor products preserve the *-structure: for *-algebras A and B, the algebraic tensor product A \otimes B becomes a *-algebra via (a \otimes b)^* = a^* \otimes b^*, yielding infinite-dimensional examples when at least one factor is, such as B(H) \otimes \mathbb{C}. This operation facilitates combining structures, like operators with polynomials, in applications to quantum mechanics.

Non-examples

Algebras lacking involution

The algebra of $2 \times 2 upper triangular matrices over \mathbb{C}, denoted UT_2(\mathbb{C}) and consisting of matrices of the form \begin{pmatrix} a & b \\ 0 & c \end{pmatrix} with a,b,c \in \mathbb{C}, provides a concrete example of an associative algebra that admits no *-involution. A *-involution on this algebra would be a conjugate-linear map ^* satisfying (xy)^* = y^* x^*, (x^*)^* = x for all x,y \in UT_2(\mathbb{C}), and preserving the algebra structure. The diagonal subalgebra, isomorphic to \mathbb{C} \oplus \mathbb{C}, must inherit the standard complex conjugation under any such involution, fixing the orthogonal idempotents E = \begin{pmatrix} 1 & 0 \\ 0 & 0 \end{pmatrix} and F = \begin{pmatrix} 0 & 0 \\ 0 & 1 \end{pmatrix} pointwise (up to the conjugation on coefficients, but as they are real in basis form). Let N = \begin{pmatrix} 0 & 1 \\ 0 & 0 \end{pmatrix} span the off-diagonal part. The relations include EN = N, NF = N, FN = NE = 0, and N^2 = 0. Applying the involution yields N^* = (EN)^* = N^* E and N^* = F N^*, alongside N^* F = 0 and E N^* = 0. Assume N^* = \begin{pmatrix} \alpha & \beta \\ 0 & \gamma \end{pmatrix} for \alpha, \beta, \gamma \in \mathbb{C}. Then N^* E = \begin{pmatrix} \alpha & 0 \\ 0 & 0 \end{pmatrix}, so N^* = N^* E implies \beta = 0 and \gamma = 0, hence N^* = \alpha E. But F N^* = 0, so N^* = F N^* implies \alpha E = 0, hence \alpha = 0 and N^* = 0. This contradicts the injectivity of the involution (as N \neq 0). Hence, no such *-involution exists. Certain division algebras also fail to admit a *-involution, particularly those lacking a compatible real form (an involution restricting to the identity on the real scalars). Examples arise in central simple algebras over number fields, such as cyclic algebras of degree greater than 2 that are isomorphic to their opposites via an anti-automorphism but possess no order-2 anti-automorphism. For instance, consider the cyclic division algebra D = (K/F, \sigma, 11), where K/F is a degree-3 Galois extension of a degree-4 extension F/\mathbb{Q} with Galois group generated by \sigma (order 3) and \tau (order 4) satisfying \tau \sigma = \sigma^{-1} \tau, and the symbol 11 ensures division over \mathbb{Q}_{11}. This D admits an anti-automorphism extending a Galois action but no involution, as its degree exceeds 2 and D \not\cong D^{\rm op} in a way compatible with order 2. Similarly, symbol algebras like (a, \omega)_{\omega, F} over \mathbb{Q}(\zeta_l) with l = \mathrm{lcm}(5,n), n=5, a = (6 + \sqrt{5})/\tau(6 + \sqrt{5}), and verification over \mathbb{Q}_{31} yield division algebras of odd degree with anti-automorphisms but no *-involution due to Galois cohomology obstructions. Higher Cayley-Dickson extensions beyond the quaternions, when adjusted to remain division (e.g., via generalized doublings over number fields), often lose the real conjugation structure, failing to admit a *-involution preserving the division property.

Incompatible structures

A common incompatibility arises when a map intended as an involution is instead an automorphism rather than an anti-automorphism. In a *-algebra, the involution must satisfy (xy)^* = y^* x^*, reversing the order of multiplication, whereas an automorphism preserves it: \phi(xy) = \phi(x) \phi(y). A classic example is the Frobenius automorphism in finite fields of characteristic p, defined by \phi(a) = a^p for a in the field \mathbb{F}_{p^n}. This map is a field automorphism, hence multiplicative in the forward order, but fails the anti-automorphism property required for a *-structure, even though commutativity makes the distinction moot in this commutative setting—it illustrates the general requirement for non-commutative algebras where order reversal is essential. Another frequent error involves maps that are anti-automorphisms but not involutive, meaning \sigma^2 \neq \mathrm{id}. For instance, certain division algebras possess anti-automorphisms without any involution existing on the algebra. In cyclic algebras D = (K/F, \sigma, u) over number fields, where K/F is a Galois extension with Galois group generated by \sigma of order n > 2 and another element \tau of order $4m \geq 4 satisfying \tau \sigma = \sigma^{-1} \tau, an anti-automorphism f can be defined by extending \tau on K and fixing u \in F^\times; however, no involution exists because D \not\cong D^{\mathrm{op}}, the opposite algebra. Specific constructions, such as n=3, m=1, and K the splitting field of x^4 + 5x + 5 = 0 and x^3 - 18x + 18 = 0 over \mathbb{Q} with u=11, yield explicit division rings of degree $3with this property. Similar examples arise in symbol algebras(a, \omega)_{n,F}forn \geq 3over\mathbb{Q}(\nu)with\nua primitivel-th root of unity (l = \mathrm{lcm}(5,n)) and achosen appropriately, where the anti-automorphism extends an automorphism\tauwith\tau(a) = a^{-1}, but again D \not\cong D^{\mathrm{op}}precludes an involution. These cases highlight how anti-automorphisms of higher order can mimic involutions superficially but violate the** = \mathrm{id}$ axiom. In skew fields (division rings), a further incompatibility occurs when the map fails conjugate-linearity over \mathbb{C}, ( \lambda a )^* = \bar{\lambda} a^*. Consider the algebra M_n(\mathbb{R}) of real n \times n matrices, viewed as an algebra over \mathbb{C}. The transpose map A \mapsto A^T is an anti-automorphism that is \mathbb{C}-linear: (\lambda A)^T = \lambda A^T for \lambda \in \mathbb{C}, since transposition does not conjugate complex scalars. This linearity over \mathbb{C} violates the antilinearity required for a *-involution, necessitating instead the conjugate transpose A \mapsto A^\dagger = \overline{A^T} to achieve conjugate-linearity. Thus, while suitable over \mathbb{R}, the transpose cannot serve as a *-involution when extending scalars to \mathbb{C}.

Extensions

Normed *-algebras

A normed *-algebra is a *-algebra A over the complex numbers equipped with a norm \|\cdot\| making (A, \|\cdot\|) a normed space (often submultiplicative: \|\mathbf{x} \mathbf{y}\| \leq \|\mathbf{x}\| \|\mathbf{y}\| for all \mathbf{x}, \mathbf{y} \in A). A -norm additionally satisfies \|\mathbf{x}^*\| = \|\mathbf{x}\| for all \mathbf{x} \in A, and a pre-C-norm further requires \|\mathbf{x}^* \mathbf{x}\| = \|\mathbf{x}\|^2. Such norms are termed -norms or pre-C-norms, distinguishing them from more general involutive norms. If A is complete with respect to this norm, it is called a Banach *-algebra. A prototypical example is the group algebra L^1(G) for a locally compact group G, where elements are equivalence classes of integrable functions on G, multiplication is given by convolution (f * g)(t) = \int_G f(s) g(s^{-1} t) \, ds, the norm is the L^1-norm \|f\|_1 = \int_G |f(s)| \, ds, and the involution is f^*(t) = \overline{f(t^{-1})}. This structure satisfies the -norm properties \|f^*\|_1 = \|f\|_1 and submultiplicativity, but not the pre-C-identity \|f^* f\|_1 = \|f\|_1^2 in general, with completeness following from the properties of the Lebesgue integral. In the context of representation theory, normed -algebras admit faithful representations on Hilbert spaces, but without the full spectral control of C-algebras; instead, their completions under a pre-C*-norm yield C*-algebras, providing a pathway to the Gelfand-Naimark framework via universal constructions. This completion process preserves the involution and algebraic operations, embedding the original structure as a dense *-subalgebra. Positive elements in a normed -algebra are typically defined within its Banach completion, where a self-adjoint element h = h^* (i.e., a Hermitian element) is positive if its spectrum \sigma(h) \subseteq [0, \infty). Equivalently, h \geq 0 iff h = \mathbf{x}^* \mathbf{x} for some \mathbf{x} \in A. This notion underpins order structures in these algebras, facilitating applications in operator theory prior to full C-completion.

C*-algebras

A C*-algebra is a complex Banach -algebra equipped with a norm satisfying the C-condition: for all elements a in the algebra, \|a^* a\| = \|a\|^2. This defining property ensures that the involution is isometric, meaning \|a^*\| = \|a\| for all a, and that the norm coincides with the spectral radius of each element. C*-algebras arise as completions of normed -algebras under a norm compatible with the C-condition. The Gelfand-Naimark theorem establishes a concrete realization for every abstract C*-algebra: it is isometrically *-isomorphic to a closed -subalgebra of the bounded linear operators on some Hilbert space. For commutative C-algebras, this representation simplifies to the algebra of continuous functions vanishing at infinity on a locally compact Hausdorff space X, denoted C_0(X). Prominent examples include the algebra of compact operators K(H) on a separable infinite-dimensional Hilbert space H, which forms a non-unital C*-algebra, and the bounded operators B(H), which is unital. Commutative instances encompass C(X) for a compact Hausdorff space X, where the involution is complex conjugation and the norm is the sup-norm. Every C*-algebra admits a faithful *-representation on a Hilbert space, constructed via the Gelfand-Naimark-Segal (GNS) procedure applied to a state (a positive linear functional of norm 1). This yields a cyclic representation where the state corresponds to the inner product with a distinguished vector, ensuring the representation preserves the algebra structure and norm.

Skew variants

In *-algebras over fields of characteristic not equal to 2, the space of elements decomposes uniquely as a direct sum of the symmetric elements S = \{ x \in A \mid x^* = x \} and the skew elements K = \{ x \in A \mid x^* = -x \}, where A = S \oplus K as vector spaces. This decomposition, often referred to as the Jordan-Lie decomposition, highlights the structural roles of these subspaces: S forms a Jordan algebra under the symmetrized product x \circ y = (xy + yx)/2, while K forms a Lie algebra under the commutator bracket [x, y] = xy - yx. A skew variant of a *-algebra arises by redefining the involution as the skew involution \sigma(x) = -x^*, which is also an involution since \sigma(\sigma(x)) = -(-x^{**}) = x. Under \sigma, the self-adjoint elements are precisely the original skew elements, satisfying \sigma(x) = x if and only if x^* = -x, thereby interchanging the roles of the symmetric and skew subspaces in the decomposition. This perspective is useful for studying structures where the skew-Hermitian part assumes the "positive" or self-adjoint role, such as in certain representations or real forms. Representative examples include the algebra of pure imaginary quaternions \operatorname{Im} \mathbb{H} = \{ ai + bj + ck \mid a,b,c \in \mathbb{R} \}, a 3-dimensional real subspace of the quaternion algebra \mathbb{H} equipped with the standard conjugation involution ^*. For any pure imaginary quaternion q, q^* = -q, so \sigma(q) = q, making all elements self-adjoint under the skew involution; moreover, \operatorname{Im} \mathbb{H} is closed under the commutator and isomorphic to the Lie algebra \mathfrak{su}(2). Similarly, \mathfrak{su}(2) itself consists of all $2 \times 2 complex traceless skew-Hermitian matrices under the adjoint involution A^* = \overline{A}^T, so the skew involution renders the entire space self-adjoint while preserving the Lie bracket structure. In the context of real forms of complex semisimple Lie algebras, skew variants appear prominently in compact real forms, where the Lie algebra is realized as the space of skew-Hermitian elements with respect to a complex structure. These forms, such as \mathfrak{su}(n) as the compact real form of \mathfrak{sl}(n, \mathbb{C}), are generated entirely by their skew-Hermitian elements under the Lie bracket, and the Cartan decomposition \mathfrak{g} = \mathfrak{k} \oplus \mathfrak{p} (with \mathfrak{k} the skew-symmetric part fixed by a Cartan involution) aligns with the skew-symmetric subspace playing a generating role. Such structures, including those related to Kac-Moody extensions, emphasize how skew elements underpin the algebra's generation in non-split real forms.

History and Applications

Historical development

The origins of *-algebras can be traced to the early 20th century, particularly the 1920s, when mathematicians like Emmy Noether and Emil Artin explored involutions within ring theory to describe real structures in fields and division algebras. Noether's work on ideals and Artin's extensions of structure theory for algebras over fields laid foundational concepts for anti-automorphisms, such as conjugation, that preserve algebraic identities and model symmetries in hypercomplex number systems. These involutions provided a framework for self-adjoint operations, influencing later abstract algebraic developments without yet incorporating the full *-algebra structure. In the 1930s and 1940s, the theory advanced significantly through operator theory on Hilbert spaces, where John von Neumann and Francis J. Murray introduced rings of operators closed under adjoints, establishing the involution as the adjoint operation to capture self-adjointness and unitarity. Their seminal 1936 paper formalized these "rings of operators," now recognized as early examples of von Neumann algebras, which are *-algebras of bounded operators weakly closed and including the identity. This work shifted focus from finite-dimensional rings to infinite-dimensional settings, emphasizing spectral theory and equivalence classes of projections to classify operator structures. Subsequent papers by Murray and von Neumann in the early 1940s further refined the role of *-involutions in resolving paradoxes in quantum mechanics and ergodic theory. Post-World War II, the formalization of -algebras crystallized with the development of C-algebras, beginning with Israel Gelfand and Mark Naimark's 1943 theorem embedding normed *-algebras into the bounded operators on Hilbert space, ensuring representation via *-homomorphisms that preserve the involution and norm. This result, published in Matematicheskii Sbornik, established the representational foundation for abstract -algebras, linking commutative cases to continuous functions on compact spaces via the Gelfand transform. Irving Segal built on this in 1947 by defining C-algebras as closed self-adjoint subalgebras of operators, introducing irreducibility criteria and the term itself in his analysis of operator representations. These contributions in the late 1940s and 1950s solidified *-algebras as a central tool in functional analysis, with ongoing refinements through the 1950s in spectral theorems and Banach algebra theory. From the 1960s onward, -algebras expanded into non-commutative geometry, pioneered by Alain Connes, who integrated them with operator algebras to model non-commutative spaces like foliations and quantum systems. Connes' work in the 1970s and 1980s, culminating in his 1994 book Noncommutative Geometry, used von Neumann and C-algebras to define spectral triples, extending differential geometry to *-algebraic settings and linking to quantum groups via Hopf algebra structures. This modern framework connected *-algebras to symmetries in quantum field theory and particle physics, with quantum groups—formalized by Vladimir Drinfeld and Michio Jimbo in 1985—providing non-commutative analogs of Lie groups represented through *-compatible bialgebras.

Key applications

-Algebras play a central role in quantum mechanics, where the self-adjoint elements of the C-algebra of bounded operators B(\mathcal{H}) on a Hilbert space \mathcal{H} represent physical observables, such as position and momentum, with the involution ensuring that observables are closed under adjoint operations to yield real-valued expectation values. This framework extends to unbounded self-adjoint operators affiliated with the algebra, allowing rigorous treatment of essential observables like the Hamiltonian, whose spectral decomposition via projection-valued measures determines energy eigenvalues and eigenstates. More abstractly, C*-algebras model quantum systems by encapsulating the algebraic relations among observables independently of the underlying Hilbert space, facilitating the study of symmetries and dynamics in systems ranging from finite-dimensional spin models to infinite-dimensional field theories. In functional analysis, the Gelfand-Naimark-Segal (GNS) construction provides a bridge to spectral theory by associating each positive linear functional (state) on a C*-algebra A with a cyclic representation \pi_\omega: A \to B(\mathcal{H}_\omega) on a Hilbert space \mathcal{H}_\omega, where the inner product is defined by \langle \pi_\omega(a) \xi_\omega, \pi_\omega(b) \xi_\omega \rangle = \omega(b^* a) for the vacuum vector \xi_\omega. This representation realizes the spectrum of self-adjoint elements in A as the support of the spectral measure, enabling the functional calculus f(a) = \int_{\sigma(a)} f(\lambda) \, dE(\lambda) for normal elements and yielding decompositions essential for analyzing operator spectra and resolvents without assuming a priori representations. Such tools underpin the classification of states and the study of ergodic actions on C*-algebras, with applications in dynamical systems where invariant measures correspond to traces on the algebra. Non-commutative geometry, as developed by Alain Connes, treats *-algebras as the coordinate algebras of "non-commutative spaces," replacing classical manifolds with spectral triples (A, \mathcal{H}, D), where A is a *-algebra, \mathcal{H} a Hilbert space representation, and D a Dirac operator—a self-adjoint unbounded operator with compact resolvent satisfying [D, a] bounded for a \in A. The Dirac operator encodes the metric geometry via the Connes distance d(p, q) = \sup \{ |\phi(p) - \phi(q)| : \|\|[D, \phi]\| \leq 1, \phi \in A \}, generalizing Riemannian distances, and facilitates differential forms and connections, as in the universal differential calculus \Omega^1(A) = A \otimes_{A^e} A / \langle da = 1 \otimes a - a \otimes 1 \rangle. This structure yields index theorems, such as pairings between K-theory classes in A and cyclic cohomology from D, recovering classical results like the Atiyah-Singer index theorem for foliations and enabling models of particle physics, including the standard model via product geometries of continuum and finite non-commutative spaces. Convolution -algebras, notably the reduced group C-algebra C_r^*(G) for a locally compact group G, form the basis for Fourier analysis on groups, where the Fourier transform maps convolution on L^1(G) to multiplication on the irreducible representations, generalizing classical Fourier transforms to non-abelian settings. In signal processing, these algebras support efficient algorithms for filtering and transforms on group-structured data, such as images invariant under group actions (e.g., rotations via the rotation group SO(2)), by decomposing signals into irreducible components via the Peter-Weyl theorem and applying group convolutions (f * g)(h) = \int_G f(k) g(k^{-1} h) \, dk for invariance-preserving operations. This approach enhances applications in image analysis and pattern recognition, where non-commutative convolutions outperform abelian ones in capturing symmetries, as demonstrated in morphological processing on finite groups for noise reduction and feature extraction.

References

  1. [1]
    [PDF] Functional Analysis and Operator Algebras - Portland State University
    May 9, 2022 · search for “universal property” in Wikipedia [52].) 4.7.1 ... (pronounced “star algebra”). If a is an element of a ∗-algebra, then a ...
  2. [2]
    star-algebra in nLab
    Apr 11, 2025 · Any plain algebra becomes a star-algebra by equipping it with the identity anti-involution. ... Wikipedia, star-algebra. Discussion for group ...
  3. [3]
    [PDF] Strongly J-Clean Rings with Involutions - arXiv
    Feb 7, 2013 · An involution of a ring R is an anti-automorphism whose square is the identity map. A ring R with involution ∗ is called a ∗-ring. All C ...
  4. [4]
    Structure of a Certain Class of Rings with Involution
    Nov 20, 2018 · Let R be a ring with involution *, and let Z denote the center of R. In R let S = {x ∈ R|x* = x} be the set of symmetric elements ...
  5. [5]
    Homomorphisms of Jordan Rings of Self-Adjoint Elements - jstor
    Let Xe denote the set of self-adjoint elements h=h*. Then 3X is a special. Jordan ring. In this paper we shall study the homomorphisms of the rings.
  6. [6]
    [PDF] C*-algebras, states and representations
    Definition 1. A Banach *-algebra A is called a C*-algebra if kx⇤xk = kxk2, x 2 A. Proposition 1. Let A be a C*-algebra. 1. kx⇤k = kxk;. 2. If A does not ...Missing: mathematics | Show results with:mathematics
  7. [7]
    [PDF] Introductory C*-algebra Theory - University of Waterloo
    Definition. A C*-algebra is a Banach space A, equipped with and as- sociative bilinear multiplcation (a, b) 7→ ab and a conjugate linear, anti-.Missing: mathematics | Show results with:mathematics
  8. [8]
    Involutions | SpringerLink
    Jun 29, 2021 · In this chapter, we define the standard involution on a quaternion algebra. In this way, we characterize division quaternion algebras as noncommutative ...
  9. [9]
    [PDF] OPERATOR ALGEBRAS: THE BANACH ALGEBRA APPROACH
    Page 1. OPERATOR ALGEBRAS: THE BANACH ALGEBRA APPROACH. Page 2. Page 3. Serban Str˘atil˘a and László Zsidó. OPERATOR ALGEBRAS: THE BANACH ALGEBRA APPROACH.
  10. [10]
    [PDF] An Introduction to Operator Algebras Laurent W. Marcoux March 30 ...
    Mar 30, 2005 · class of operator algebras. We shall divide our analysis into the ... Finally, a Banach *-algebra is an involutive Banach algebra A whose.
  11. [11]
    None
    Summary of each segment:
  12. [12]
    Mathematical foundations of quantum mechanics : Von Neumann ...
    Jul 2, 2019 · Mathematical foundations of quantum mechanics. by: Von Neumann, John, 1903-1957. Publication date: 1955. Topics: Matrix mechanics. Publisher ...
  13. [13]
    [PDF] involutions on rank 16 central simple algebras
    An involution a is split if and only if D has a-invariant quaternion subalgebras. A natural question arises as to when an involution o on D splits. Examples of ...
  14. [14]
    [PDF] Quaternion algebras - John Voight
    Mar 20, 2025 · Moreover, quaternions often encapsulate unique features that are absent from the general theory (even as they provide motivation for it). With ...
  15. [15]
    [math/0612103] Positive polynomials in scalar and matrix variables ...
    Dec 4, 2006 · ... *-algebra, the appropriate setting for analyzing inequalities on polynomials having matrix variables. We sketch some of these developments ...
  16. [16]
    [0910.2210] Tensor Products of Operator Systems - arXiv
    Oct 12, 2009 · ... *-algebra yet has the property that for every C*-algebra A, the minimal and maximal tensor product of S and A are equal. Comments: 41 pages ...
  17. [17]
    [PDF] Division algebras with an anti-automorphism but no involution - CSUN
    Thus, the existence of a division ring with an anti-automorphism but no involution gives examples of projective geometries with dualities but no polarities. All ...
  18. [18]
    Frobenius homomorphism - PlanetMath
    Mar 22, 2013 · is a field homomorphism, called the Frobenius homomorphism, or simply the Frobenius map on F F . ... , and is called the Frobenius automorphism ...<|separator|>
  19. [19]
    Conjugate transpose - StatLect
    Discover what the operation of conjugate transposition does on a matrix, what symbols are used to denote it, and what properties it has.
  20. [20]
    The group of automorphisms and anti ... - MathOverflow
    Mar 8, 2017 · My question: Is it true that G is the free amalgamated product of two of its subgroups, namely, the affine and the triangular automorphisms and ...
  21. [21]
    [PDF] Introduction to Normed ∗-Algebras and their Representations, 6th ...
    Jul 26, 2008 · Definition (algebra). An algebra is a complex vector space A equipped with an associative multiplication (a, b) 7→ ab satisfying the.
  22. [22]
    Banach Algebras and the General Theory of *-Algebras
    This is the first volume of a two volume set that provides a modern account of basic Banach algebra theory including all known results on general Banach ...
  23. [23]
    [PDF] NOTES ON C∗-ALGEBRAS Contents 1. A first look at C
    A useful source of examples and motivation for C∗-theory are the group C∗-algebras. Indeed, one can view a group C∗-algebra as encoding the (infinite- ...
  24. [24]
    The Embedding Problem in Algebras with Involution - ResearchGate
    Nov 11, 2024 · Thus, a+a∗is symmetric and a−a∗is skew-symmetric for any a∈A. Therefore, we have a decomposition A=A ...
  25. [25]
    [PDF] A Taste of Jordan Algebras - Galileo and Einstein
    It is divided into eight sections: the origin of the species, the genus of related Jordan structures, and links to six other areas of mathematics (Lie algebras ...
  26. [26]
    [PDF] Lie Theory as Illustrated by SU(2)
    More formally, when we use the quaternionic view, we can view. R. 3 = Rhii ⊕ Rhji ⊕ Rhki as the purely imaginary quaternions. Then, taking t ∈ S3,q ∈ R3 ...
  27. [27]
    [PDF] Representations of su(2) 1 Lie and linear groups
    hence the Lie algebra su(2) of SU(2) consists of all traceless two-by-two skew-hermitian matrices: su(2) = {X ∈ Mat(2, C) : X = −X†, tr X = 0}. A basis for ...
  28. [28]
    [PDF] Classification of Real Forms of Semisimple Lie Algebras
    So let us list real forms of simple Lie algebras that we know so far. 1. Type An−1. We have the split form sln(R), the compact form su(n), and ...
  29. [29]
    [PDF] cartan and iwasawa decompositions in lie theory - UBC Math
    of structure in its Lie algebra. We start with a Lie algebra L of matrices and seek a decomposition into symmetric and skew-symmetric parts. To get this ...
  30. [30]
    [PDF] The Book of Involutions Max-Albert Knus Alexander Merkurjev ...
    To achieve this, Albert developed a theory of central simple algebras with involution, based on the theory of simple algebras initiated a few years earlier.
  31. [31]
    On the imbedding of normed rings into the ring of operators ... - EuDML
    Гельфанд, И.М., and Наймарк, MA. "On the imbedding of normed rings into the ring of operators in Hilbert space." Matematiceskij sbornik 54.2 (1943): 197-217.
  32. [32]
    [PDF] Noncommutative Geometry Alain Connes
    This book is the English version of the French “Géométrie non commutative” pub- lished by InterEditions Paris (1990). After the initial translation by S.K. ...
  33. [33]
    [PDF] Operator Algebras and Unbounded Self-Adjoint Operators
    Jun 16, 2015 · C∗-algebras were first considered in quantum mechanics to build a model of physical observables. In particular, one can describe a physi- cal ...
  34. [34]
    [PDF] operator algebras and their application in physics - arXiv
    Aug 22, 2022 · C‹-algebra of continuous functions on some compact Hausdorff space. Now we consider a non-commutative example of a C‹-algebra. Example 1.10.
  35. [35]
    [PDF] The Gelfand-Naimark-Segal construction
    Dec 22, 2017 · Our goal is to show that every C∗-algebra is isometrically ∗-isomorphic to a ∗-subalgebra of bounded linear operators on some Hilbert space.
  36. [36]
    Group Algebras in Signal and Image Processing - ScienceDirect
    The multiplication in group algebra is called “group convolution.” The selection of coefficients from the cyclic convolution to generate output for the linear ...
  37. [37]
    Group morphology with convolution algebras - ACM Digital Library
    The proposed computational approach is based on group convolution algebras, which extend classical convolutions and the Fourier transform to non-commutative ...