Fact-checked by Grok 2 weeks ago

Jordan algebra

A Jordan algebra is a vector space over a (typically or complex numbers) equipped with a commutative bilinear , denoted x \circ y, that satisfies the Jordan identity (x^2 \circ y) \circ x = x^2 \circ (y \circ x), where x^2 = x \circ x. This structure generalizes the algebra of operators in , originating from work by , , and in 1934 to model observables as Hermitian matrices while avoiding non-commutativity issues in associative algebras. Key properties of Jordan algebras include commutativity (x \circ y = y \circ x) and power-associativity, meaning that powers of a single element associate, though the algebra as a whole is non-associative. Many algebras are quadratic, defined via a linear U_x(y) = x \circ y \circ x that satisfies the fundamental identity U_{x \circ y \circ x} = U_x U_y U_x, enabling a unified treatment through quadratic forms rather than just bilinear products. They often admit a unit element and are nondegenerate, meaning no nonzero element annihilates the entire under the U-operator. Special Jordan algebras arise from associative algebras by symmetrizing the product (x \circ y = \frac{1}{2}(xy + yx)), while exceptional ones, such as the 27-dimensional Albert algebra over the octonions, cannot be embedded this way. Finite-dimensional simple Jordan algebras over algebraically closed fields of characteristic not 2 are classified by the Renaissance Structure Theorem as either the one-dimensional algebra over the field, spin factors (related to quadratic forms), Hermitian matrices over split composition algebras, or the exceptional 27-dimensional Albert algebra. For formally real (Euclidean) Jordan algebras, the –Wigner theorem states they decompose into direct sums of simple ideals, including algebras of Hermitian matrices over the reals, complexes, or quaternions, the spin factor, and the Albert algebra H_3(\mathbb{O}). Zelmanov's solution to the exceptional problem confirms that the only simple exceptional finite-dimensional Jordan algebras are the Albert algebras. Structural tools like the Peirce decomposition, which splits the algebra relative to an idempotent into eigenspaces J_2 \oplus J_1 \oplus J_0, aid in analyzing subalgebras and representations. Jordan algebras have profound applications beyond algebra, including connections to exceptional groups and algebras via the Tits–Koecher construction, which embeds them into algebras of types E_6, E_7, and F_4. They underpin the geometry of Hermitian symmetric spaces and bounded symmetric domains in , as well as projective geometries like Moufang planes through coordinatization theorems. In , the characterizes Jordan C*-algebras as special or exceptional 27-dimensional ones, linking to operator algebras. Their study continues to influence , quadratic forms, and even extensions.

Fundamentals

Definition

A Jordan algebra is defined as a J over a \Phi, typically the real numbers \mathbb{R} or complex numbers \mathbb{C}, equipped with a bilinear multiplication operation \circ: J \times J \to J that is commutative, meaning x \circ y = y \circ x for all x, y \in J. This operation, often denoted by the circle product x \circ y, provides a structure that generalizes aspects of associative algebras while relaxing the associativity condition. Unlike associative algebras, where the multiplication satisfies (x y) z = x (y z), Jordan algebras are inherently non-associative, allowing for more flexible algebraic structures that arise naturally in certain physical and geometric contexts. The bilinearity ensures that the multiplication distributes over addition and is homogeneous with respect to scalar multiplication from \Phi, i.e., (a x) \circ y = x \circ (a y) = a (x \circ y) for a \in \Phi. In many cases, especially when derived from an underlying associative algebra with multiplication xy, the Jordan product is given explicitly by the symmetrized form x \circ y = \frac{1}{2}(xy + yx), which enforces commutativity. A minimal example illustrating this structure is the algebra of $1 \times 1 matrices over \Phi, which is isomorphic to \Phi itself with the standard multiplication serving as the Jordan product; here, the operation is both commutative and associative, but the framework accommodates broader non-associative cases. The complete axiomatic characterization, including the Jordan identity that ensures a form of modified associativity, is addressed in subsequent sections on axioms and properties.

Axioms and Properties

A Jordan algebra over a field of characteristic not 2 is a vector space equipped with a bilinear operation ∘, called the Jordan product, satisfying two fundamental axioms: commutativity and the Jordan identity. The commutativity axiom states that for all elements x, y in the algebra, x \circ y = y \circ x. This ensures the product is symmetric, distinguishing Jordan algebras from non-commutative structures like Lie algebras. The Jordan identity, which encodes a weakened form of associativity particularly for squares, is given by (x^2 \circ y) \circ x = x^2 \circ (y \circ x), where x^2 = x \circ x. This axiom guarantees "associativity in the square," meaning that the algebra behaves associatively when one factor is a square, facilitating the study of quadratic structures central to the theory. From these axioms, several key structural properties follow. One important consequence is power-associativity, which asserts that powers of a single element associate: for any natural number n, x^{2n+1} = x \circ x^{2n}, where powers are defined inductively via the Jordan product (e.g., x^3 = x \circ (x \circ x)). More generally, x^m \circ x^k = x^{m+k} for positive integers m, k. To sketch the derivation, first note that commutativity implies x \circ x^2 = x^2 \circ x = x^3. Applying the Jordan identity with y = x yields (x \circ x) \circ x^2 = x \circ (x \circ x^2), simplifying to x^2 \circ x^2 = x \circ x^3 = x^4, establishing x^4 = (x^2)^2. Induction on even powers proceeds similarly, and linearizing the Jordan identity or using the associated operator U_x(y) = 2(x \circ y) \circ x - x^2 \circ y (which satisfies U_x(x) = x^3) confirms U_{x^2}(x) = x^5, and iteratively U_{x^{2n}}(x) = x^{2n+1}, yielding the result. This property ensures that the subalgebra generated by a single element is associative, simplifying computations and classifications. Associated with the Jordan product is a quadratic form Q: V \to V defined by Q(x) = x^2. This form inherits homogeneity from bilinearity: for any scalar a, Q(ax) = (ax)^2 = a^2 x^2 = a^2 Q(x). Moreover, the polarization identity allows recovery of the product via x \circ y = \frac{1}{2} \left( Q(x+y) - Q(x) - Q(y) \right), confirming that the entire structure is determined by the quadratic form, with the Jordan axioms imposing conditions on Q such as Q(x) \circ y = 2x \circ (x \circ y) - x^2 \circ y. These features make Jordan algebras quadratic in nature, linking them to and optimization. A Jordan algebra is unital if it admits an e satisfying x \circ e = x for all x. In this case, the quadratic form simplifies further, as U_e = \mathrm{id} and Q(e) = e^2 = e, enabling like the Peirce decomposition relative to idempotents. Not all Jordan algebras are unital, but every Jordan algebra embeds into a unital one, often called its unital hull.

Historical Development

Origins in Physics

The origins of Jordan algebras trace back to efforts in the early 1930s to provide an algebraic framework for the observables in , particularly those represented by operators on . , motivated by the need to formalize the symmetric aspects of quantum measurements, proposed in 1933 a non-associative product for observables using the anticommutator defined as \{x, y\} = \frac{1}{2}(xy + yx). This operation ensures symmetry, reflecting the expectation values of sums and products of observables, while avoiding the full associativity of classical algebras to accommodate quantum non-commutativity. In 1933–1934, Jordan collaborated with and to extend this idea into a broader suitable for . Their joint work culminated in the seminal paper "On an Algebraic Generalization of the Quantum Mechanical Formalism," where they introduced what are now called Jordan-von Neumann-Wigner algebras, specifically tailored to the algebra of symmetric observables such as Hermitian matrices. These algebras were designed to capture the statistical properties of quantum measurements in a coordinate-free manner, emphasizing the role of the Jordan product in defining addition and for observables. The framework allowed for a generalization beyond standard , potentially applicable to more complex physical systems. In this work, they also classified all finite-dimensional simple formally real Jordan algebras over the reals, showing they are direct sums of five types: algebras of Hermitian matrices over the reals, complexes, or quaternions; the 27-dimensional exceptional Albert algebra; and spin factors. A key innovation was linking the Jordan product to Lie algebras via the commutator [x, y] = xy - yx, which encodes the antisymmetric, dynamical aspects of quantum , while the Jordan product handled the symmetric, observational side. This duality provided a unified algebraic view of quantum symmetries and . However, the non-associativity of the Jordan product posed early challenges, as it was introduced precisely to circumvent paradoxes arising from the non-commutativity in quantum processes, such as the dependence on measurement order, yet it complicated the derivation of consistent physical laws without additional axioms.

Mathematical Formalization

The formal definition of Jordan algebras as an algebraic structure was introduced by Pascual Jordan, John von Neumann, and Eugene Wigner in their 1934 paper, where they axiomatized the non-associative multiplication for observables in quantum mechanics using the symmetrized product x \circ y = \frac{1}{2}(xy + yx) to ensure commutativity while allowing non-associativity. This axiomatization emphasized a unital vector space over a field equipped with a commutative bilinear product satisfying the Jordan identity (x^2 \circ y) \circ x = x^2 \circ (y \circ x), decoupling the structure from physical interpretations and establishing it as a purely mathematical entity. The term "Jordan algebra" was coined by A. Adrian Albert in 1946. In the 1930s and 1940s, Nathan Jacobson played a pivotal role in developing the algebraic theory, extending the classification to semi-simple finite-dimensional algebras over algebraically closed fields of characteristic not 2, where he showed such algebras decompose into direct sums of simple ideals, with simple ones arising from associative algebras via symmetrization or exceptional constructions. Jacobson's work included introducing U-operators U_x y = (xy + yx)/2 and cubic norm forms, laying groundwork for coordinatization theorems that linked algebras to algebras of dimensions 1, 2, 4, or 8. The 1950s saw further advancements in structure theory by A. Adrian Albert and Tonny A. Springer. Albert introduced the Peirce decomposition and proved the exceptionality of the Albert algebra H_3(\mathbb{O}), while Springer developed the cubic norm construction and linked Jordan algebras to algebraic groups and octonion planes, culminating in theorems on and homogeneity. In the , the theory recognized deeper connections to s and quadratic modules, with Kevin McCrimmon introducing quadratic Jordan algebras, where the structure derives from a Q(x) and to a B(x,y) = Q(x+y) - Q(x) - Q(y), unifying the original axiomatic approach with geometric and norm-based interpretations. This quadratic formulation facilitated links to homogeneous cones and Tits constructions, enhancing the algebraic framework without altering the core classification. As of 2025, the classification of finite-dimensional simple algebras remains complete, with no major gaps since Zel'manov's 1979 confirmation that they fall into classical families (Hermitian, quadratic form-based) or the exceptional 27-dimensional type over division rings of characteristic not 2 or 3; however, continues on infinite-dimensional cases and representations over rings.

Examples

Finite-Dimensional Examples

One prominent class of finite-dimensional Jordan algebras consists of the spaces of Hermitian matrices over division rings of finite dimension over the reals. For the real numbers, the Jordan algebra \mathfrak{h}_n(\mathbb{R}) is formed by the n \times n real symmetric matrices with the Jordan product A \circ B = \frac{1}{2}(AB + BA); this algebra has dimension \frac{n(n+1)}{2} and n. Over the complex numbers, the analogous \mathfrak{h}_n(\mathbb{C}) uses n \times n complex Hermitian matrices with the same product formula, yielding dimension n^2 and n. For the quaternions, \mathfrak{h}_n(\mathbb{H}) comprises n \times n quaternionic Hermitian matrices under the identical product, with dimension n(2n-1) and n. Spin factor algebras provide another fundamental family of finite-dimensional examples. These are realized as the \mathbb{R}^{2k-1} \oplus \mathbb{R} (for integer k \geq 2) or more generally \mathbb{R}^m \oplus \mathbb{R} (with m \geq 2), equipped with the product (x, \alpha) \circ (y, \beta) = (\alpha y + \beta x, x \cdot y + \alpha \beta), where \cdot denotes the standard inner product on \mathbb{R}^m. Such an algebra has dimension m+1 and rank 2. The finite-dimensional Jordan algebras over the reals—specifically, the formally real ones—are completely classified into the matrix algebras \mathfrak{h}_n(\mathbb{K}) for \mathbb{K} = \mathbb{R}, \mathbb{C}, \mathbb{H} and n \geq 3; the spin factors of dimension at least 3; and the exceptional Albert algebra of dimension 27 and rank 3, realized as the Jordan algebra of $3 \times 3 Hermitian matrices over the with the symmetrized product. Every finite-dimensional formally real Jordan algebra is an orthogonal of simple ideals from this .

Exceptional Jordan Algebras

The exceptional Jordan algebra, commonly referred to as the Albert algebra, is a 27-dimensional Jordan algebra consisting of all Hermitian matrices over the Cayley-Dickson \mathbb{O}. Its Jordan product is defined by A \circ B = \frac{1}{2}(AB + BA), where AB denotes the usual matrix product with entries multiplied using the non-associative octonion multiplication. This structure was first constructed by A. A. Albert in 1950 as an example of a non-special Jordan algebra, highlighting its departure from those derivable from associative algebras. The algebra is equipped with a cubic norm form N(A) = \det(A), which is preserved under the Jordan product and plays a central role in its properties. A general element of the Albert algebra can be written as \begin{pmatrix} a_{11} & x_{12} & x_{13} \\ \overline{x_{12}} & a_{22} & x_{23} \\ \overline{x_{13}} & \overline{x_{23}} & a_{33} \end{pmatrix}, where a_{11}, a_{22}, a_{33} \in \mathbb{R} and x_{12}, x_{13}, x_{23} \in \mathbb{O}. This yields the dimension $3 + 3 \times 8 = 27, as each off-diagonal octonionic entry contributes 8 real dimensions, with the Hermitian condition ensuring the lower triangle is determined by conjugation. A standard basis consists of the three diagonal matrix units e_{ii} (i=1,2,3) with real coefficients, together with basis matrices for the off-diagonal blocks corresponding to the standard basis \{1, e_1, \dots, e_7\} of \mathbb{O}, where e_0 = 1 is the real unit and \{e_1, \dots, e_7\} span the pure imaginary octonions. The multiplication in the Albert algebra inherits the non-associativity of the , particularly in the off-diagonal entries. For instance, the product of two basis elements involving off-diagonal parts follows the octonion multiplication table: if e_i \cdot e_j = -\delta_{ij} + \sum_k c_{ijk} e_k (with c_{ijk} the of \mathbb{O}, such as e_1 \cdot e_2 = e_3, e_2 \cdot e_3 = e_1, etc., up to the Fano plane associations), then the (i,j)-entry of A \circ B includes terms like \frac{1}{2}(a_{ik} b_{kj} + b_{ik} a_{kj}) symmetrized over octonion products. This sketch illustrates how the exceptional nature arises from the alternativity of , ensuring the Jordan axioms hold despite non-associativity. The Albert algebra is unique as the only exceptional simple finite-dimensional Jordan algebra over algebraically closed fields of characteristic zero, distinguishing it from the classical types derived from associative structures. Its construction emerges from the Freudenthal-Tits magic square, a tableau associating pairs of normed division algebras (reals, complexes, quaternions, ) to Lie algebras and Jordan algebras, where the ( \mathbb{O}, \mathbb{O} ) entry yields this 27-dimensional exceptional case. The algebra has rank 3, meaning the maximum of a generic element (or the dimension of the span of its minimal idempotents) is 3. In the Peirce decomposition relative to a primitive idempotent e (satisfying e \circ e = e and generating a 3-dimensional ), the Albert algebra decomposes as J = J_1(e) \oplus J_{1/2}(e) \oplus J_0(e), where J_1(e) = \mathbb{R} e (1-dimensional), J_{1/2}(e) \cong \mathbb{O} (8-dimensional, capturing the octonionic action), and J_0(e) is 18-dimensional, reflecting the perpendicular complement. This decomposition underscores the octonionic genesis and is generic for primitive idempotents in the algebra. The structure group of the Albert algebra, which preserves the generic cubic norm, is the exceptional Lie group of type E_6(-78), of dimension 78. This group encodes the symmetries of the algebra beyond its inner derivations, linking it to broader exceptional structures in .

Special Classes

Hermitian Jordan Algebras

Hermitian Jordan algebras arise naturally in the context of and , where they model the of observables. Consider a complex \mathcal{H} equipped with a conjugation (an antilinear ), and let \mathfrak{h} denote the real of operators on \mathcal{H} with respect to this conjugation. The Jordan product is defined by a \circ b = \frac{1}{2}(ab + ba) for a, b \in \mathfrak{h}, where ab is the operator composition. This structure satisfies the Jordan axioms of commutativity a \circ b = b \circ a and the Jordan identity (a^2 \circ b) \circ a = a^2 \circ (b \circ a), where a^2 = a \circ a, forming a Hermitian Jordan over the reals. More abstractly, Hermitian Jordan algebras can be constructed from an associative algebra A over \mathbb{R} or \mathbb{C} equipped with an \ast (an anti-automorphism satisfying (xy)^\ast = y^\ast x^\ast and x^{\ast\ast} = x). The space H(A, \ast) = \{ x \in A \mid x^\ast = x \} of fixed points under \ast inherits the Jordan product x \circ y = \frac{1}{2}(xy + yx), yielding a unital commutative algebra over \mathbb{R}. In the finite-dimensional case, typical examples include spaces of Hermitian matrices H_n(\mathbb{D}) over a \mathbb{D}. These constructions capture both finite- and infinite-dimensional settings, with the realization providing a infinite-dimensional . A fundamental property is the equivalence between Hermitian Jordan algebras and Euclidean Jordan algebras over the reals. A Euclidean Jordan algebra is a finite-dimensional formally real Jordan algebra (where \sum x_i^2 = 0 implies each x_i = 0) equipped with a positive definite inner product compatible with the Jordan structure, admitting a and a proper open of positive elements. Every finite-dimensional Hermitian Jordan algebra, when viewed over \mathbb{R}, is Euclidean, and conversely, the Euclidean ones admit a Hermitian realization via an appropriate on an enveloping . This equivalence underpins their role in symmetric cone analysis and optimization. The classification of finite-dimensional simple Hermitian Jordan algebras is given by the Jordan-von Neumann-Wigner theorem: every such algebra is isomorphic to one of the following matrix types over the real division algebras \mathbb{R}, \mathbb{C}, \mathbb{H} (quaternions), or \mathbb{O} (octonions)—specifically, the Hermitian matrices H_n(\mathbb{R}) for any n \geq 1, H_n(\mathbb{C}) for n \geq 1, H_n(\mathbb{H}) for $1 \leq n \leq 3, or H_3(\mathbb{O}) (the exceptional 27-dimensional Albert algebra)—or to a spin factor JSpin_n, which is the Jordan algebra associated to a quadratic form of signature (n, 2). This classification highlights their connection to classical groups and exceptional Lie structures, with the octonionic case marking the boundary of associativity. Spectral decomposition is a hallmark feature, generalizing the diagonalization of Hermitian matrices. For any element x in a finite-dimensional Hermitian Jordan algebra J, there exists a spectral resolution x = \sum_{i=1}^r \lambda_i e_i, where the \lambda_i \in \mathbb{R} are distinct real eigenvalues, the e_i are pairwise orthogonal primitive idempotents (projections satisfying e_i^2 = e_i and e_i \circ e_j = 0 for i \neq j), and \sum e_i is the unit element. The rank of J equals the maximum number of such eigenvalues over all elements, determining the degree of the generic minimal polynomial \det(\lambda I - L_x) = \prod (\lambda - \lambda_i). This decomposition facilitates the study of automorphisms and isotopies. The positive elements in a Hermitian Jordan algebra form an open regular convex cone C_+(J), consisting of those x for which the spectrum \operatorname{Spec}(x) \subseteq \mathbb{R}_{>0}. This cone coincides with the interior of the cone of squares \{ y^2 \mid y \in J \} and is self-dual with respect to the trace inner product \langle x, y \rangle = \operatorname{Tr}(L_x L_y), where L_x y = x \circ y. Positive elements generate the order structure, enabling applications in and symmetric cone domains, where the cone's homogeneity ensures unique solvability of certain equations. Hermitian Jordan algebras are formally real, as their positive cone prevents nontrivial sums of squares from vanishing.

Formally Real Jordan Algebras

A formally real Jordan algebra over the real numbers is defined as a Jordan algebra J in which the only solution to the equation \sum_{i=1}^n x_i^2 = 0 (with x_i \in J) is x_1 = \cdots = x_n = 0 for every finite n. This condition, known as formal reality, ensures that no nontrivial of squares vanishes, distinguishing these algebras from those over complex fields where such relations may exist. Formal reality implies the existence of a natural partial on J, making it an . Specifically, the positive C = \{ x \in J \mid \operatorname{spec}(x) \subseteq \mathbb{R}_{\geq 0} \} (where \operatorname{spec}(x) denotes the spectrum of the multiplication by x) is a proper open , and x \leq y if y - x \in C. This is compatible with the Jordan product and turns J into an ordered space where sums of squares are positive elements. The Koecher-Vinberg theorem establishes a between finite-dimensional simple formally real Jordan algebras and homogeneous self-dual cones in . Under this correspondence, the algebra J is equipped with an inner product making the trace form \sigma(x, y) = \operatorname{tr}(L_{x \circ y}) (where L_z is left multiplication by z) positive definite, and the associated cone C is self-dual with respect to \sigma, meaning C = \{ y \in J \mid \sigma(x, y) \geq 0 \ \forall x \in C \}. Moreover, C is homogeneous, as the structure group of J (automorphisms preserving the order) acts transitively on its interior. Prominent examples of such cones include the cone of symmetric matrices S_n^+(\mathbb{R}), arising from the Jordan algebra of n \times n real symmetric matrices with product x \circ y = \frac{1}{2}(xy + yx), and the second-order (or ice-cream) cone \{ (t, x) \in \mathbb{R} \times \mathbb{R}^n \mid t \geq \|x\|_2 \}, associated with the spin factor algebra \mathbb{R} \oplus \mathbb{R}^n with elements (t, v) and product (t, v) \circ (s, w) = (ts + \langle v, w \rangle, tw + sv), where \langle \cdot, \cdot \rangle is the standard inner product. The associated cone C in a formally real Jordan algebra is Archimedean, meaning that if n x \leq u for all natural numbers n (with order unit u), then x \leq 0; this prevents infinitesimal elements and ensures the order is compatible with the induced by the inner product. The facial structure of C is governed by the of elements: for x = \sum \lambda_i e_i with orthogonal primitive idempotents e_i and distinct eigenvalues \lambda_i, faces correspond to subsets of the eigenspaces, yielding a where exposed faces are principal (generated by minimal idempotents) and the face lattice mirrors the algebra's Peirce decomposition J = \bigoplus J_{ij}.

Structure Theory

Derivations

In a Jordan algebra J over a of characteristic not equal to 2, a is a D: J \to J satisfying the Leibniz rule with respect to the Jordan product \circ: D(x \circ y) = D(x) \circ y + x \circ D(y) for all x, y \in J. This condition ensures that derivations preserve the , acting as automorphisms. The set of all derivations, denoted \mathrm{Der}(J), forms a under the bracket [D_1, D_2] = D_1 D_2 - D_2 D_1. Inner derivations arise from the algebra itself and are generated by commutators of left multiplications. Specifically, for x, y \in J, the map D_{x,y}: J \to J defined by D_{x,y}(z) = x \circ (y \circ z) - y \circ (x \circ z) = [L_x, L_y](z), where L_x(z) = x \circ z, is an inner derivation. The span of all such D_{x,y} forms the ideal of inner derivations \mathrm{Inn}(J) in \mathrm{Der}(J). In semi-simple Jordan algebras, every derivation is inner, meaning \mathrm{Der}(J) = \mathrm{Inn}(J). For finite-dimensional simple Jordan algebras over algebraically closed fields of characteristic zero, the derivation algebra \mathrm{Der}(J) is a . This simplicity reflects the rigid structure of such algebras, with \mathrm{Der}(J) serving as the of the \mathrm{Aut}(J). For classical simple examples, such as the algebra J_n^+(A) of n \times n Hermitian matrices over a A with (with n \geq 3), \mathrm{Der}(J) is isomorphic to the special linear \mathfrak{sl}(n, A) or its variants adjusted for the involution. In the exceptional case of the 27-dimensional Albert algebra J_3(\mathbb{O}) over the , \mathrm{Der}(J) is the 52-dimensional of type F_4. These structures highlight how derivations capture the symmetries of Jordan algebras, with proofs often relying on tools like the Peirce decomposition to analyze representations. Outer derivations, which would quotient to \mathrm{Der}(J)/\mathrm{Inn}(J), vanish in these simple finite-dimensional cases, implying complete reducibility of derivation modules and reinforcing the absence of non-trivial extensions.

Peirce Decomposition

In a Jordan algebra J over a field of characteristic not 2, an idempotent element e \in J satisfies e \circ e = e. Relative to such an e, the Peirce decomposition provides a direct sum decomposition of the algebra into eigenspaces of the left multiplication operator L_e(x) = e \circ x: J = V_2(e) \oplus V_1(e) \oplus V_0(e), where V_k(e) = \{ x \in J \mid e \circ x = \frac{k}{2} x \} for k = 0, 1, 2. These subspaces are orthogonal with respect to the Jordan product in the sense that V_2(e) \circ V_0(e) = 0, and the decomposition is unique, with every element x \in J expressible as x = x_2 + x_1 + x_0 where x_k \in V_k(e). The multiplication rules in the Peirce decomposition are governed by the Jordan identity and symmetry, yielding: \begin{align*} V_2(e) \circ V_2(e) &\subseteq V_2(e), \\ V_2(e) \circ V_1(e) &\subseteq V_1(e), \\ V_1(e) \circ V_1(e) &\subseteq V_2(e) \oplus V_0(e), \\ V_1(e) \circ V_0(e) &\subseteq V_1(e), \\ V_0(e) \circ V_0(e) &\subseteq V_0(e). \end{align*} These rules imply that V_2(e) and V_0(e) behave as subalgebras, while V_1(e) acts as a bimodule over them, facilitating the analysis of . In the more general setting of multiple orthogonal idempotents e_1, \dots, e_r with \sum e_i = 1 (a complete orthogonal ), the Peirce decomposition refines to a J = \bigoplus_{i,j=1}^r V_{ij}, where V_{ij} = \{ x \in J \mid e_k \circ x = \frac{\delta_{ki} + \delta_{kj}}{2} x \ \forall k \}, and the product satisfies V_{ij} \circ V_{kl} \subseteq V_{il} \oplus V_{kj}. Orthogonal idempotents e and f (satisfying e \circ f = 0) exhibit complete orthogonality in their joint Peirce spaces, meaning V_i(e) \circ V_j(f) = 0 unless i = j = 2 or specific connecting indices. Primitive idempotents, defined as those that cannot be expressed as a nontrivial sum of orthogonal nonzero idempotents, play a key role in this : for a primitive e, the space V_1(e) = 0, simplifying J = V_2(e) \oplus V_0(e) with V_2(e) = \mathbb{F} e (a one-dimensional subalgebra). The rank of a unital Jordan algebra J, denoted \operatorname{rank}(J), is the maximal number of pairwise orthogonal primitive idempotents summing to the unit element, and a complete orthogonal system of primitive idempotents induces a Peirce decomposition that fully coordinatizes J. In simple finite-dimensional Jordan algebras over algebraically closed fields, the Peirce spaces relative to a idempotent determine the algebra's type. Specifically, the structure of V_1(e) as a bimodule over the division algebra components of V_2(e) and V_0(e) classifies J as a Hermitian Jordan algebra J_n(D, \Gamma) or the exceptional Albert algebra, with the and multiplication in these spaces fixing the parameters n and the composition D. For instance, if \dim V_1(e) = 2 \dim V_2(e) - 1, the algebra is of spin type, distinguishing it from matrix-like or exceptional forms.

Structure Groups

The \operatorname{Aut}(J) of a Jordan algebra J consists of all bijective linear maps T: J \to J that preserve the Jordan product, satisfying T(x \circ y) = T(x) \circ T(y) for all x, y \in J. This group captures the symmetries of the algebra while fixing the multiplicative structure exactly, and it forms a when J is finite-dimensional over \mathbb{R} or \mathbb{C}, with its given by the derivations of J. The structure group \operatorname{Str}(J) generalizes the by allowing transformations that preserve the quadratic structure up to scaling. Specifically, it comprises all invertible linear maps T \in \mathrm{GL}(J) such that Q(Tx) = \lambda(T) Q(x) for some scalar \lambda(T) \in F^\times, where Q(x) = x^2 is the associated to the Jordan product, and T(x) \circ T(y) = \lambda(T) T(x \circ y) for all x, y \in J. Equivalently, T preserves the U-operators via U_{T(x)}(Ty) = \lambda(T) T U_x(y), ensuring T induces an from J to a scaled of itself. The map \lambda: \operatorname{Str}(J) \to F^\times is a , and \operatorname{Aut}(J) embeds as the where \lambda(T) = 1. In the case of Euclidean Jordan algebras, equipped with a positive definite inner product \langle x, y \rangle = \operatorname{tr}(x \circ y), the structure group coincides with the , consisting of transformations T satisfying \langle Tx, Ty \rangle = \lambda(T) \langle x, y \rangle for all x, y \in J. This group acts transitively on the interior of the positive cone and plays a central role in the of symmetric cones associated to J. The reduced structure group \operatorname{Str}_0(J) is the kernel of the scalar homomorphism \lambda, comprising those T \in \operatorname{Str}(J) with \lambda(T) = 1, thus preserving the quadratic form exactly without scaling. For unital Jordan algebras, this includes transformations fixing the unit element, T(1) = 1. The of the structure group decomposes as \operatorname{Strl}(J) = L(J) \oplus \operatorname{Der}(J), where \operatorname{Der}(J) is the algebra and L(J) consists of the "linear" elements generated by commutators involving U-operators of invertible elements. This decomposition highlights how infinitesimal structure transformations combine multiplicative symmetries (automorphisms and derivations) with adjustments from the .

Generalizations

Infinite-Dimensional Jordan Algebras

Infinite-dimensional Jordan algebras extend the axiomatic definition of finite-dimensional ones to vector spaces of arbitrary dimension, typically equipped with a complete norm to form Banach spaces. A Jordan-Banach algebra, or JB-algebra, is a unital Banach space over the reals that is a Jordan algebra satisfying the Jordan identity x \circ (x^2 \circ y) = x^2 \circ (x \circ y) for all elements x, y, along with the norm conditions \|x^2\| = \|x\|^2 and \|x \circ y\| \leq \|x\| \|y\| for all x, y. These algebras are formally real, meaning that if a finite sum of squares equals zero, each term is zero. Unlike finite-dimensional cases, the norm ensures completeness and compatibility with the algebraic structure, addressing challenges like unbounded operators in Hilbert spaces. The U-operator is defined as U_x(y) = 2 (x \circ y) \circ x - x^2 \circ y, where x^2 = x \circ x. A prominent example of an infinite-dimensional JB-algebra arises from the self-adjoint elements of a . For a unital \mathcal{A}, the space \mathcal{A}_{sa} of self-adjoint elements forms a JB-algebra under the Jordan product a \circ b = \frac{1}{2}(ab + ba) and the operator norm inherited from \mathcal{A}. This construction captures many physically relevant structures, such as the observables in quantum mechanics modeled on B(H)_{sa}, the self-adjoint bounded operators on a Hilbert space H. Spin factor JB-algebras, generalizing finite-dimensional versions, also provide infinite-dimensional instances, consisting of sequences with a quadratic form and Clifford multiplication. These examples illustrate how JB-algebras embed into operator-theoretic settings without requiring associativity. In contrast to the finite-dimensional case, where simple Jordan algebras are classified into five types plus the exceptional 27-dimensional Albert algebra, no complete classification exists for infinite-dimensional JB-algebras. Efforts focus on structural decompositions rather than exhaustive lists, with JB-algebras often analyzed via their ideals, state spaces, or representations. A key result, the Gelfand-Naimark theorem for Jordan algebras, states that every unital JB-algebra possesses a unique closed exceptional ideal J, which is either zero or finite-dimensional and isomorphic to the Albert algebra, such that the quotient A/J is isometrically isomorphic to a closed Jordan subalgebra of B(H)_{sa} for some H. This implies that infinite-dimensional JB-algebras are "essentially special," embeddable into associative structures modulo a finite exceptional kernel, highlighting the predominance of operator-like behavior in high dimensions. As of 2025, open problems persist in the structure theory of JB-algebras, particularly for non-type I or non-nuclear variants, where full analogs of finite-dimensional tools like Peirce decompositions and groups remain incomplete. Challenges include characterizing multiplier algebras and deriving explicit representations without relying on embeddings for all cases, underscoring the ongoing development of non-commutative in these settings.

Jordan Operator Algebras

Jordan operator algebras arise in the context of embedding algebras into the framework of on s. Specifically, JB-algebras are defined as the norm-closed real Jordan subalgebras of the operators on a , B(H)_{\mathrm{sa}}, equipped with the \|\cdot\| and the symmetrized product a \circ b = \frac{1}{2}(ab + ba). Abstractly, a JB-algebra is a Banach space over the reals that is also a Jordan algebra satisfying the conditions \|x^2\| = \|x\|^2 for all x and \|U_x(y)\| \leq \|x\| \|y\| for all x, y, where U_x(y) = 2 (x \circ y) \circ x - x^2 \circ y is the Jordan multiplication . These structures capture the non-associative aspects of quantum observables while inheriting analytic properties from Banach spaces. A fundamental result, known as the Jordan Gelfand-Naimark theorem, establishes that every JB-algebra admits a faithful isometric Jordan homomorphism onto a norm-closed Jordan subalgebra of B(H)_{\mathrm{sa}} for some H, up to a unique closed exceptional ideal J such that the quotient A/J embeds concretely. This representation highlights the operator-theoretic nature of JB-algebras. Additionally, by a theorem of Kadison, every JB-algebra (modulo its exceptional part) arises isometrically as the self-adjoint elements of a under the Jordan product. Exceptional JB-algebras, such as the 27-dimensional Albert algebra of Hermitian 3×3 matrices over the , do not embed non-trivially into B(H)_{\mathrm{sa}} and are rare, occurring only in finite dimensions; no infinite-dimensional exceptional JB-algebras exist. JBW-algebras extend JB-algebras to the von Neumann algebraic setting: they are the weak operator topology (w.o.t.)-closed Jordan subalgebras of B(H)_{\mathrm{sa}}, or equivalently, the weak*-closed JB-algebras with a predual. These algebras possess rich , including orthomodular lattices of projections and normal states, analogous to algebras. Key properties include unitization: every JB- or JBW-algebra admits a unique (up to ) unitization \tilde{A}, obtained by adjoining an if necessary. The center Z(A) of a JBW-algebra A, consisting of elements commuting with all others under the Jordan product, is itself a JBW-algebra and plays a role in decomposing A into central factors. JBW-algebras admit a type decomposition into direct sums of factors of types I, II_1, II_\infty, and III, mirroring the Murray- classification for von Neumann algebras, based on the behavior of projections and traces. Type I factors correspond to B(H)_{\mathrm{sa}} or algebras over rings, while types II and III involve projections without minimal ones. This classification facilitates the study of structural groups and automorphisms in infinite dimensions.

Quadratic Jordan Algebras

Quadratic Jordan algebras provide an axiomatic framework for Jordan algebras that emphasizes forms and operators, allowing treatment over rings of arbitrary without relying on a linear bilinear product. Introduced by Kevin McCrimmon, this formulation defines the structure through a quadratic map U: J \to \mathrm{End}(J) on a J over a unital commutative ring \Phi, where U is quadratic in its argument, meaning U_{\lambda x + \mu y} = \lambda^2 U_x + \mu^2 U_y + \lambda \mu T(x,y) for a bilinear map T, and satisfies key identities: U_{U_x(y)} = U_x U_y U_x (the fundamental identity) and V_{x,y} U_x = U_x V_{y,x}, where V_{y,z} w = (y \circ z \circ w) + (w \circ z \circ y) is the linearized triple product (once polarized). This setup ensures the algebra is unital with U_1 = \mathrm{id}_J, and the quadratic nature captures the Jordan product implicitly through squaring: x^2 = U_x(1). The Jordan product x \circ y is recovered via polarization of the quadratic form associated to U. Specifically, the trace form T(x) = U_x(1) and the bilinear polarization Q(x,y) = \frac{1}{2} (Q(x+y) - Q(x) - Q(y)), where Q(x) = \frac{1}{2} T(x^2), yield x \circ y = \frac{1}{2} (T(x)y + T(y)x - Q(x,y) \cdot 1). The trilinear T-operator further encodes the structure: T(x,y)z = 2(x \circ y \circ z) + 2(z \circ y \circ x) - x \circ (y^2 \circ z) - z \circ (y^2 \circ x), allowing full recovery of the product from this operator alone in certain cases. This equivalence holds because quadratic Jordan algebras correspond precisely to quadratic modules over \Phi-modules with compatible endomorphisms satisfying the above identities, providing a module-theoretic perspective on the algebraic structure. A Jordan algebra is termed a T-algebra if its product is fully determined by the bilinear T-, meaning all operations derive from T(x,y) without additional structure. All finite-dimensional Jordan algebras over fields are T-algebras, as their structure groups and Peirce decompositions ensure the T- suffices to reconstruct the algebra uniquely; for example, the simple finite-dimensional cases—such as Hermitian matrices over division algebras or spin factors—are all T-algebras. In contrast, infinite-dimensional Jordan algebras may fail to be T-algebras, as embeddings into associative structures can introduce dependencies not captured solely by T, though nondegenerate cases often retain T-algebra properties. Quadratic Jordan algebras relate to associative algebras through symmetrization: given an associative algebra A with unit, the symmetrized Jordan algebra A^+ has product x \circ y = \frac{1}{2}(xy + yx), and its quadratic operator is U_x(y) = x y x when restricted to elements where this holds. Over rings containing \frac{1}{2}, every quadratic Jordan algebra is isomorphic to such a symmetrized subalgebra of an associative envelope, linking the quadratic formulation to the classical linear one via this construction.

Jordan Superalgebras and J-Structures

Jordan superalgebras extend the theory of Jordan algebras to incorporate a \mathbb{Z}_2-grading, consisting of a vector space J = J_{\bar{0}} \oplus J_{\bar{1}} over a field of characteristic not 2, equipped with an even bilinear product that preserves the grading and satisfies supercommutativity xy = (-1)^{|x||y|} yx. The even part J_{\bar{0}} forms an associative algebra, while the odd part J_{\bar{1}} is a bimodule over J_{\bar{0}} via the action x \cdot y = xy = yx for x \in J_{\bar{0}}, y \in J_{\bar{1}}. The structure satisfies the super-Jordan identity (a \cdot b) \cdot (c \cdot d) + (-1)^{|b||c|} (a \cdot c) \cdot (b \cdot d) + (-1)^{|b||d| + |c||d|} (a \cdot d) \cdot (b \cdot c) = ((a \cdot b) \cdot c) \cdot d + (-1)^{|c||d| + |b||c|} ((a \cdot d) \cdot c) \cdot b + (-1)^{|a||b| + |a||c| + |a||d| + |c||d|} ((b \cdot d) \cdot c) \cdot a. A J-structure provides a unified framework for superalgebras, extending the original concept introduced by to the super setting by incorporating a compatible bracket on the graded space. This bracket allows the Jordan product to be recovered from the structure, such as via symmetrized operations, and facilitates of derivations, automorphisms, and related Lie-theoretic in the presence of grading and sign rules. Representative examples include Jordan superalgebras arising from Clifford modules, where the even part is the associative Jordan algebra of the associated to a on a , and the odd part consists of the module with the product defined via the Clifford action to satisfy the super identities. Another key example is the Jordan superalgebra derived from the orthosymplectic \mathfrak{osp}(m|2n), whose even part is the associative algebra of symmetric elements and odd part the skew-symmetric elements, structured to form a unital simple Jordan superalgebra. In these structures, the even part's associativity ensures embeddability into associative superalgebras for special cases, while the odd part functions as a faithful bimodule, supporting the overall supercommutativity and identity satisfaction. Exceptional Jordan superalgebras, not embeddable in this way, exhibit more complex behaviors but share the graded properties. Brief generalizations to color gradings replace the \mathbb{Z}_2-grading with \mathbb{Z}^k- or gradings, introducing color-dependent commutation factors and analogous identities, though these remain less developed. As of 2025, Jordan superalgebras appear in supersymmetric theories to model graded symmetries and operators, but their full classification, especially for infinite-dimensional instances, remains incomplete despite advances in finite-dimensional cases over algebraically closed fields.

Applications

In Symmetric Spaces

Jordan algebras play a fundamental role in the classification of irreducible symmetric spaces of non-compact type through their connection to Cartan decompositions and associated triple systems. In a symmetric space G/K, where G is a semisimple and K is a maximal compact subgroup, the Lie algebra \mathfrak{g} admits a Cartan decomposition \mathfrak{g} = \mathfrak{k} \oplus \mathfrak{p}, with \mathfrak{k} the Lie algebra of K and \mathfrak{p} the orthogonal complement identified with the tangent space at the base point. The space \mathfrak{p} inherits a canonical structure of a Jordan triple system via the triple product \{x, y, z\}_{\mathfrak{p}} = [[x, y], z], where [ \cdot, \cdot ] denotes the Lie bracket, enabling the geometric realization of symmetric spaces in terms of Jordan structures. Euclidean Jordan algebras, which are formally real Jordan algebras equipped with a positive definite inner product compatible with the Jordan product, parametrize the irreducible Hermitian symmetric spaces of tube type. These spaces are realized as tube domains V + i \Omega, where V is the underlying real of the Jordan algebra and \Omega is the associated symmetric cone of squares of positive elements. The classification of such spaces corresponds directly to the simple Euclidean Jordan algebras, including the algebras of symmetric matrices over \mathbb{R}, Hermitian matrices over \mathbb{C} or \mathbb{H}, and the exceptional Albert algebra over the . A key result in this context is the theorem of Loos establishing a bijective correspondence between irreducible symmetric cones and formally real Jordan algebras. Specifically, every irreducible symmetric cone arises as the cone of positive invertible elements in a Euclidean Jordan algebra, and conversely, every such algebra determines a unique irreducible symmetric cone via its interior of squares, which carries the structure of a Riemannian symmetric space. A representative example is the upper half-space, associated to the symmetric space \mathrm{Sp}(2n, \mathbb{R})/\mathrm{U}(n), which corresponds to the Jordan algebra of n \times n real symmetric positive definite matrices under the Jordan product x \circ y = \frac{1}{2}(xy + yx). This space is the tube domain over the cone of positive definite matrices, illustrating how the Jordan structure governs the geometry of the symmetric space. In these symmetric spaces, components are analyzed using the structure of the symmetric cones, where faces correspond to ideals in the Jordan algebra. The Shilov boundary, the minimal set in the topological maximizing the Bergman , coincides with the set of tripotents of maximal in the associated Jordan triple system, providing a compactification and facilitating on the space.

In Quantum Mechanics and Physics

In noncommutative geometry, the Jordan product provides a framework for describing observables in , particularly through the symmetrized product on self-adjoint elements of C*-algebras, which forms a Jordan algebra. This approach aligns with ' spectral triples, where the real structure encodes physical observables like masses and mixing angles in the , reviving Jordan structures as a tool for nonassociative coordinate algebras compatible with Majorana fermions. Specifically, the self-adjoint part of the algebra captures measurable quantities, allowing distances in the Connes sense to be defined via Jordan operators, thus bridging with observables. In quantum information theory, formally real Jordan algebras underpin the structure of quantum states and entanglement detection through their self-dual positive cones, which characterize the base of the state space in finite-dimensional systems. Developments in the and leveraged these cones to define entanglement witnesses as elements violating separability conditions in the Jordan framework, enabling operational reconstructions of quantum theory from algebraic axioms. For instance, the Jordan product distinguishes quantumness from classicality by identifying states where the product fails to preserve classical probabilistic structures, with applications to multipartite entanglement in composite systems modeled as tensor products of simple Jordan algebras. The 27-dimensional exceptional Jordan algebra over the , known as the Albert algebra, yields the exceptional groups F_4 as its and E_6 as the reduced structure group, playing a role in compactifications and heterotic models. In , this algebra parameterizes the of Calabi-Yau manifolds or vertex operators in E_6 representations, linking algebraic symmetries to unification schemes beyond the . These structures appear in the large-N limit of Jordan matrix models, connecting to on exceptional backgrounds. In , Jordan triples derived from degree-three Jordan algebras model generalized spacetimes with Freudenthal-Tits magic squares, where the triple product defines and symmetry in higher-dimensional geometries. Murat Günaydin's work constructs phase spaces and models from these algebras, incorporating exceptional groups to describe horizons and solutions as symmetric spaces. This framework extends to non-associative settings, with the Jordan structure preserving Lorentz invariance in curved backgrounds.

References

  1. [1]
    [PDF] A Taste of Jordan Algebras - Galileo and Einstein
    This book covers Jordan algebras, their structure, and their role in mathematics, including their origin in quantum physics and the 27-dimensional Albert  ...
  2. [2]
    Jordan Algebras and their applications - Project Euclid
    In this article I want to sketch for nonexperts what Jordan algebras are why people might want to study such strange objects. I start with the.
  3. [3]
  4. [4]
    On an Algebraic Generalization of the Quantum Mechanical Formalism
    JORDAN, J. v. NEUMANN AND E. WIGNER. (The left sides are symmetric in a and ,3 while the right sides are not. Hence, for a > /3, we obtain valid formulas by ...
  5. [5]
    [PDF] Non-Associative Algebras and Quantum Physics - arXiv
    Sep 11, 2019 · Early attempts first by Pascual Jordan and then by Jordan, John von Neumann and Eu- gene Wigner to generalize the formalism of quantum mechanics ...
  6. [6]
    The Representation of Abstract Integrals on JSTOR
    **Summary of Main Contribution:**
  7. [7]
    Quadratic Jordan Algebras and Cubing Operations - jstor
    QUADRATIC JORDAN ALGEBRAS AND CUBING OPERATIONS BY KEVIN McCRIMMON. Abstract. In this paper we show how the Jordan structure can be derived from the squaring ...
  8. [8]
    [PDF] Lecture 21 - Jordan Algebras and Projective Spaces - Penn Math
    Apr 15, 2013 · ... dimensional formally real special Jordan algebras are. • The n × n real symmetric matrices hn(R). • The n × n complex Hermitian matrices hn(C).
  9. [9]
    A Note on the Exceptional Jordan Algebra - PNAS
    12 Whiting, Anna R., Ibid., 97, 210 (1949). A NOTE ON THE EXCEPTIONAL JORDAN ALGEBRA. BY A. A. ALBERT. DEPARTMENT OF MATHEMATICS, THE UNIVERSITY OF CHICAGO.
  10. [10]
    A Note on the Exceptional Jordan Algebra - PNAS
    A Note on the Exceptional Jordan Algebra. A. A. AlbertAuthors Info & Affiliations. July 15, 1950. 36 (7) 372-374. https://doi.org/10.1073/pnas.36.7.372.Missing: paper | Show results with:paper
  11. [11]
    [PDF] The characteristic equation of the exceptional Jordan algebra
    Abstract. The exceptional Jordan algebra [also known as the Albert algebra] is the finite dimensional algebra of 3x3 Hermitean matrices with octonionic entries.
  12. [12]
    Albert algebra in nLab
    Jul 21, 2020 · The octonionic and split-octonionic Albert algebras are (up to isomorphism) the only simple finite-dimensional formally real Jordan algebras ...Definition · Properties · Relation to 10d super-Spacetime · Automorphisms and...
  13. [13]
    [PDF] Albert algebras
    A cubic Jordan algebra J over k is said to be reduced if it isomorphic to Her3(C, Γ) for some composition k-algebra C and some diagonal matrix Γ ∈ GL3(k). Thus, ...
  14. [14]
    [PDF] Gradings on the octonions, the Albert algebra, and exceptional ...
    The gradings on the octonions and on the Albert algebra are instrumental in obtaining the gradings on the exceptional simple. Lie algebras. Page 6. Outline.
  15. [15]
    The Exceptional Simple Jordan Algebras - jstor
    while the theory of simple Jordan algebras corresponds to the original work in 1934 of Jordan, von Neumann, and Wigner ([6]) and Albert ([1]): all simple ...
  16. [16]
    [math/0203010] Magic squares and matrix models of Lie algebras
    We review the Tits-Freudenthal construction of the magic square, which includes the exceptional Lie algebras as the octonionic case of a construction in terms ...
  17. [17]
    [PDF] Exceptional Jordan Algebras
    If J is an Albert algebra, Str(J ) = Instr(J ) is the norm preserving group and is of type E6. Michel Racine (Ottawa). Exceptional Jordan Algebras. March,2013.
  18. [18]
    Structurable Algebras and Groups of Type E6 and E7 - ScienceDirect
    It is well known that every group of type F4 is the automorphism group of an exceptional Jordan algebra, and that up to isogeny all groups of type 1E6 with ...
  19. [19]
    [PDF] derivations of jordan c*-algebras
    Further, a JW-factor of dimension +3,4,6 is a spin factor if and only if it is not reversible [26; Th. 7.1]. The following theorem is the main result of this ...
  20. [20]
    None
    ### Summary of Formally Real Jordan Algebras from https://www2.gwu.edu/~spwm/Jordan%20Algebras.pdf
  21. [21]
    Infinite dimensional Jordan algebras and symmetric cones
    Dec 1, 2017 · A celebrated result of Koecher and Vinberg asserts the one-one correspondence between the finite dimensional formally real Jordan algebras and ...
  22. [22]
    [PDF] PoS(CORFU2019)163 - Proceeding of science
    We shall, by contrast, base our treatment on an appropriate finite dimensional Jordan algebra2 suited for a quantum theory - permitting, in particular, a ...
  23. [23]
    [PDF] Order structures, Jordan algebras, and geometry
    This is a 27- dimensional formally real Jordan algebra which is also known as the Albert algebra. In all but the spin factors the Jordan prod- uct is as in ...<|control11|><|separator|>
  24. [24]
    [PDF] arXiv:1707.03610v1 [math.RA] 12 Jul 2017
    Jul 12, 2017 · This correspondence is furnished by the assertion that given a finite dimensional formally real Jordan algebra A, which is a Hilbert space in ...
  25. [25]
    [PDF] The Jordan Algebraic Structure of the Circular Cone
    The most well-known examples of symmetric cones are the nonnegative orthant cone in the space of real numbers R, the second-order cone in the real vector space ...
  26. [26]
    [PDF] Facial structure of copositive and completely positive cones over a ...
    Feb 6, 2025 · As mentioned above, symmetric cones and Euclidean Jordan algebras have a one- to-one correspondence. For any symmetric cone K in a finite- ...
  27. [27]
    [PDF] Derivation Algebras and Multiplication Algebras of Semi-Simple ...
    In this note we investigate the Lie algebra of derivations and the Lie algebra V generated by the multiplications in any semi-simple Jordan algebra (with a.
  28. [28]
    Derivations of Jordan algebras - MSP
    A derivation of a Jordan algebra J into a module M is a linear map D of J into ikf satisfying. A definition of inner derivation of any non-associative ...
  29. [29]
    the exceptional simple lie algebras f4 and e6 - jstor
    THEOREM. The exceptional simple Lie algebra F4 of dimension 52 and rank 4 over K is the derivation algebra Z of the exceptional Jordan algebra a.
  30. [30]
    [PDF] An order-theoretic characterization of JB-algebras - arXiv
    Jul 15, 2025 · July 15, 2025. Abstract. We give an order-theoretic characterization of the JB-algebras among the complete order unit.Missing: reality | Show results with:reality
  31. [31]
    [PDF] Structure of JB*-Triples - 2. Gelfand Naimark theorem
    Theorem 2.1. 2 (Gelfand-Naimark theorem for Jordan algebras). If A is a JB- algebra, then there is unique closed ideal J such that A/J is isometrically isomor- ...<|control11|><|separator|>
  32. [32]
    [PDF] jb algebras with an exceptional ideal
    Theorem 2 implies in particular that any essentially exceptional JB algebra is of type I≤3. There is a different way to view theorem 2. Let ✓ be an ...
  33. [33]
    THE GELFAND-.NAIMARKTHEOREM - Project Euclid
    THE GELFAND-NAIMARK THEOREM FOR JB*-TRIPLES. 141 by 3 self-adjoint matrices and has a natural Jordan algebra structure. C consists of all by 2 matrices and ...
  34. [34]
    On the structure group of an infinite dimensional JB-algebra
    May 15, 2023 · Consider V = B ( H ) s a , the self-adjoint operators of a complex separable Hilbert space H with the Jordan product A ∘ B = 1 / 2 ( A B + B A ) ...
  35. [35]
    [PDF] arXiv:2206.09208v3 [math.DG] 3 Sep 2024
    Sep 3, 2024 · We endow the Banach-Lie structure group Str(V ) of an infinite di- mensional JB-algebra V with a left-invariant connection and Finsler metric, ...
  36. [36]
    Jordan operator algebras: basic theory - Blecher - Wiley Online Library
    Jan 22, 2018 · Jordan operator algebras are norm-closed spaces of operators on a Hilbert space which are closed under the Jordan product.
  37. [37]
    OPERATOR MEANS IN JB-ALGEBRAS - ScienceDirect.com
    A Jordan algebra A over real numbers is a vector space A over ℝ equipped with a bilinear product ∘ that satisfies the following identities: a ∘ b = b ∘ a , ( a ...
  38. [38]
    A Gelfand-Neumark theorem for Jordan algebras - ScienceDirect.com
    It is shown that A possesses a unique norm closed Jordan ideal J such that has a faithful representation as a Jordan algebra of self-adjoint operators on a ...
  39. [39]
    Orientation in operator algebras - PMC - PubMed Central
    By a theorem of Kadison (7) the ordering and the norm of a C*-algebra (or equivalently, its state space) determine the symmmetrized (Jordan) product ½(ab + ba) ...
  40. [40]
    [PDF] THEORY OF JORDAN OPERATOR ALGEBRAS AND OPERATOR
    A Jordan C∗-algebra is said to be a JC*-algebra if it is isometrically ∗-isomorphic to a norm-closed, Jordan ∗-subalgebra of B(H) for some Hilbert space H. A JC ...Missing: sa | Show results with:sa
  41. [41]
    [PDF] A Taste of Jordan Algebras
    This book tells the story of one aspect of Jordan structure theory: the origin of the theory in an attempt by quantum physicists to find algebraic systems more ...<|control11|><|separator|>
  42. [42]
    [PDF] Lectures On Quadratic Jordan Algebras
    Φ-module together with a product xy which is Φ-bilinear and associa- ... is the symmetric bilinear form associated with Q. Hence xyx = −yx2 +. Q(x,y) ...
  43. [43]
    Jordan superalgebra in nLab
    Feb 16, 2023 · A Jordan superalgebra is a supercommutative superalgebra with underlying ℤ / 2 -graded algebra J = J 0 ⊕ J 1 J = J_0 \oplus J_1 , where: J 0 is ...
  44. [44]
    [PDF] Simplicity of Jordan Superalgebras and Relations with Lie Structures
    In what follows algebra will mean algebra over a field F. Definition 1.1. A superalgebra A = A¯0 + A¯1 is a Z2-graded algebra, that is, A¯ ...
  45. [45]
    J-structures | SpringerLink
    Cite this chapter. Springer, T.A. (1998). J-structures. In: Jordan Algebras and Algebraic Groups. Classics in Mathematics. Springer, Berlin, Heidelberg.
  46. [46]
    Jordan Superalgebras Defined by Brackets - ResearchGate
    PDF | Jordan superalgebras defined by brackets on associative commutative superalgebras are studied. It is proved that any such superalgebra is imbedded.
  47. [47]
    [PDF] Classification of three-dimensional Jordan superalgebras - arXiv
    Aug 24, 2017 · The goal of this paper is to classify Jordan superalgebras of dimension up to three over an algebraically closed field of characteristic ...
  48. [48]
    Simple Jordan superalgebras with the even parts of Clifford type
    Mar 11, 2025 · The purpose of this paper is a partial progress towards classification of simple infinite dimensional Jordan superalgebras. First, we prove that ...
  49. [49]
    [PDF] BIMODULES OVER SIMPLE FINITE-DIMENSIONAL JORDAN ...
    is a Jordan superalgebra of a superform. Let osp(1,2) denote the Lie subsuperalgebra of M1,2(F) which consists of skewsymmetric elements with respect.
  50. [50]
    Representations of simple Jordan superalgebras - ScienceDirect.com
    Aug 26, 2020 · This paper completes description of categories of representations of finite-dimensional simple unital Jordan superalgebras over algebraically closed field of ...
  51. [51]
    [PDF] Basics on Hermitian Symmetric Spaces
    The first section explains the notion of a Cartan decomposition of a Lie algebra, which is generalization of decomposition of g = k+p. The second and the ...
  52. [52]
    [PDF] Jordan Algebras and Symmetric Manifolds
    Dec 6, 2021 · Conversely, given a complex Jordan triple V , Kaup. [6] has shown that one can construct a real Lie algebra L with decomposition L = k ⊕ p, and ...
  53. [53]
    Jordan Algebras and Harmonic Analysis on Symmetric Spaces - jstor
    ... algebra (of finite dimen- sion) making Aut(D) into a real Lie group. The symmetry - idD at 0 e D induces a Cartan decomposition g= (B. (1.2) into the Lie ...
  54. [54]
    [PDF] Geometry of the Shilov Boundary of a Bounded Symmetric Domain
    The characterization of the Shilov boundary of a bounded symmetric domain is specially nice in the approach through PHJTS, as the elements of the Shilov ...
  55. [55]
    [PDF] The geometry of physical observables - arXiv
    Mar 27, 2020 · Rethinking connes' approach to the standard model of particle physics via non-commutative geometry. New Journal of Physics, 17(2):023021 ...
  56. [56]
    [PDF] Composites and Categories of Euclidean Jordan Algebras
    Nov 8, 2020 · classes of finite dimensional simple euclidean Jordan algebras: the matrix algebras. Rn,Cn,Qn, the exceptional Jordan algebra M3(O)sa, and ...
  57. [57]
    [PDF] Defining quantumness via the Jordan product - arXiv
    Sep 18, 2013 · We propose alternative definitions of classical states and quantumness witnesses by focusing on the algebra of observables of the system. A ...
  58. [58]
    Physics The Exceptional Jordan Algebra and the Superstring
    As we have already mentioned, they can be regarded as a twenty-seven dimensional representation of E6 and so might be represented as vertex operators defined ...
  59. [59]
    [PDF] E6, Strings, Branes, and the Standard Model - Algebraic Reality
    Apr 13, 2004 · E6 corresponds to 26-dim String Theory, related to traceless J3(O)o and the symmetric space E6/F4. • E7 corresponds to 27-dim M-Theory, related ...
  60. [60]
    [PDF] arXiv:hep-th/0506010v3 26 Aug 2005
    Abstract. We study the symmetries of generalized spacetimes and correspond- ing phase spaces defined by Jordan algebras of degree three. The.
  61. [61]
    [PDF] Stabilizer Formalism for Operator Algebra Quantum Error Correction
    Feb 21, 2024 · We introduce a stabilizer formalism for the general quantum error correction framework called operator algebra quantum error correction ...Missing: JB | Show results with:JB