Fact-checked by Grok 2 weeks ago

Noncommutative geometry

Noncommutative geometry is a branch of that generalizes classical and by replacing commutative algebras of functions on spaces—such as the smooth functions C^\infty(M) on a manifold M—with noncommutative operator algebras, enabling the study of "quantum spaces" and singular structures like foliations or Penrose tilings through tools from and . This framework, primarily developed by in the 1980s and 1990s, draws inspiration from where observables like position and momentum do not commute ([q, p] = i\hbar), and it aims to unify geometric invariants with physical models by extending concepts like differential forms, metrics, and index theorems to noncommutative settings. At its core, noncommutative geometry relies on spectral triples (\mathcal{A}, \mathcal{H}, D), where \mathcal{A} is a noncommutative -algebra acting on a Hilbert space \mathcal{H}, and D is an unbounded self-adjoint operator (often a Dirac operator) that encodes the metric and differential structure of the underlying "space." In the commutative limit, when \mathcal{A} = C^\infty(M) for a compact Riemannian manifold (M, g), the spectral triple recovers classical geometry, including geodesic distances via the formula d(x,y) = \sup \{ |f(x) - f(y)| : \| [D, f] \| \leq 1, f \in \mathcal{A} \} and volume forms through residues of traces like the Dixmier trace. Key tools include cyclic cohomology, which generalizes de Rham cohomology and pairs with K-theory to yield Chern characters and index pairings, as in the noncommutative Atiyah-Singer theorem; Fredholm modules, which quantify topological invariants via the index of projections; and C-algebras with their K-groups, providing noncommutative analogues of topological invariants for spaces like the noncommutative torus A_\theta. These structures allow the treatment of pathological spaces, such as the leaf spaces of foliations via groupoid C*-algebras C^*(V, F), where classical measure theory fails due to non-Hausdorff topology. Historically, noncommutative geometry traces its origins to Werner Heisenberg's 1925 , where noncommuting matrices replaced classical Poisson brackets, and was formalized through the development of operator algebras by in the 1930s, including the Gel'fand-Naimark theorem equating commutative C*-algebras to continuous functions on compact spaces. advanced the field in the 1970s–1980s by integrating Tomita-Takesaki modular theory for traces and states, Masaki Kashiwara's KK-theory for asymptotic morphisms, and his own cyclic (developed with Daniel Kastler and Boris Tsygan), culminating in the 1994 Noncommutative Geometry. Contributions from Jean Bellissard on noncommutative solenoids and Joseph Wodzicki on residue traces further enriched the toolkit, bridging , , and analysis. In physics, noncommutative geometry has profound applications, modeling the Standard Model of particle physics as a finite noncommutative space adjoined to spacetime, where the gauge group U(1) \times SU(2) \times SU(3) emerges from the internal degrees of freedom of a spectral triple, reproducing the Glashow-Weinberg-Salam action and fermion masses via the spectral action functional \operatorname{Tr} f(D/\Lambda). It describes the quantum Hall effect through the noncommutative torus, where Hall conductivity is quantized as a Chern number in K-theory, aligning with experimental integer and fractional values. Other notable uses include noncommutative deformations of spacetimes in string theory and M-theory compactifications on tori, where duality symmetries are captured by Morita equivalence of algebras, and quantum field theories on noncommutative backgrounds to resolve ultraviolet divergences. These applications highlight noncommutative geometry's role in bridging quantum field theory, general relativity, and condensed matter physics, with ongoing research exploring almost-commutative geometries for grand unification.

Introduction

Definition and Basic Concepts

Noncommutative geometry is a mathematical framework that generalizes classical geometry by replacing commutative algebras of functions on spaces with noncommutative algebras, allowing the study of geometric structures where traditional notions of points and coordinates fail to apply directly. In classical geometry, the structure of a space, such as a compact Hausdorff topological space X, is encoded by the commutative C*-algebra C(X) of continuous complex-valued functions on X, where multiplication corresponds to pointwise operations. Noncommutativity emerges naturally in quantum mechanics, where observables are represented by operators on a Hilbert space whose multiplication is noncommutative, leading to algebras like the bounded operators on Hilbert space that capture "quantum spaces" without underlying classical points. The Gelfand-Naimark theorem establishes a foundational duality for the commutative case: every unital commutative A is isometrically *-isomorphic to C(X) for some compact X, where X is the of A, constructed via the Gelfand transform that maps elements of A to functions on its space. In the noncommutative setting, this theorem generalizes by embedding any into the algebra of bounded operators on a , interpreting noncommutative algebras as algebras of "functions" on hypothetical noncommutative spaces devoid of points, where geometric invariants are recovered through and representations. A basic example illustrating this paradigm is the M_n(\mathbb{C}) of n \times n matrices, which serves as a finite-dimensional noncommutative analog of a point: while a single point in classical corresponds to the one-dimensional \mathbb{C}, M_n(\mathbb{C}) approximates more structures, such as finite approximations to continuous spaces, and exhibits noncommutativity in its operator . The key principle of noncommutative geometry is thus the reconstruction of geometric data—such as distances, measures, and topologies—from purely algebraic information provided by the spectrum, , and unitary representations of these algebras, extending classical tools like differential forms and to the noncommutative realm. This approach relies on operator algebras, particularly C*-algebras, as the primary objects encoding noncommutative spaces.

Historical Development

The foundations of noncommutative geometry trace back to the mid-20th century, with Israel Gelfand's program in the 1940s establishing the duality between commutative Banach algebras and topological spaces via the Gelfand transform, which laid the groundwork for extending geometric ideas to noncommutative settings. This commutative framework was complemented by the 1943 work of Gelfand and Mark Naimark, who defined C*-algebras and proved their representation as operators on Hilbert spaces, providing a amenable to . In 1949, Richard Arens extended representation theorems to certain noncommutative Banach algebras, further bridging and by allowing approximations and representations that generalized classical results. The 1970s marked a pivotal shift with ' contributions to the classification of algebras, where he introduced rigidity properties and injectivity criteria that revealed deep connections between operator algebras and . Connes extended this to in 1978, associating a von Neumann algebra to any smooth of a manifold and showing that its weights correspond to operator-valued transverse measures, thus encoding geometric invariants in algebraic terms. Building on these ideas, Connes developed cyclic cohomology in the early 1980s as a noncommutative analogue of , providing tools to compute characteristic classes for foliations and leading to the initial formulation of noncommutative geometry. This culminated in the 1990 French publication of Géométrie non commutative (English edition 1994), which synthesized these developments into a comprehensive framework. In the 1990s, Connes introduced the spectral triple—a triple consisting of an , a , and a —as a central construct in his 1994 book, enabling the unification of with through noncommutative Riemannian structures and local index formulas. This publication served as a for the field, outlining how operator algebras could replace manifolds in quantum contexts while preserving metric and topological data. Concurrently, advanced noncommutative projective geometry in his 1988 Montreal lectures, proposing quantum analogues of projective spaces via noncommutative coordinate rings, which influenced subsequent algebraic developments. The 2000s and saw expansions through Kontsevich's 1997 proof of the existence of deformation quantizations for any , providing a formal star product that deforms commutative algebras into noncommutative ones while preserving brackets, thus linking to . In the , advances in noncommutative mirror symmetry, such as Alexander Polishchuk's work on homological equivalences for elliptic curves via Fukaya categories and matrix factorizations, have further integrated noncommutative tools with dualities. During the , applications emerged in , particularly through noncommutative deformations in the AdS/CFT correspondence, where spectral triples model holographic dualities between gauge theories and gravitational backgrounds on noncommutative spaces.

Motivations

From Mathematical Physics

In quantum mechanics, the Heisenberg uncertainty principle arises from the noncommutativity of and operators, fundamentally altering the classical notion of . This noncommutativity, where the commutator [x, p] = iℏ, implies that precise simultaneous measurements of position and momentum are impossible, motivating a geometric framework where space itself becomes noncommutative. Early formulations, such as Weyl quantization in the , mapped classical functions on to noncommuting operators, laying the groundwork for noncommutative geometry as a tool to describe quantum spaces. Applications in and further highlight this motivation. The noncommutative torus, introduced by Marc Rieffel in 1980, models the by treating the electron's phase space as a noncommutative , as applied by Jean Bellissard and collaborators, capturing quantized conductance plateaus observed experimentally. Similarly, the deforms the classical pointwise multiplication of functions on into a noncommutative star product, enabling quantum field theories on deformed spaces that incorporate noncommutativity at high energies or strong magnetic fields. In , noncommutative geometry provides a unified description of the through the Connes-Lott model of 1990, which reconstructs the gauge group SU(3) × SU(2) × U(1) from a finite noncommutative space with algebra C \oplus \mathbb{H} \oplus M_3(\mathbb{C}). This approach treats fermions and bosons on an equal footing, deriving the Higgs sector from the geometry of a product space involving classical and finite noncommutative components. Recent developments in the 2020s extend these ideas to , where noncommutativity resolves singularities and modifies . Using the spectral action functional, calculations yield corrections to the Bekenstein-Hawking , incorporating noncommutative effects that alter thermodynamic laws near horizons. These models suggest fuzziness at the Planck scale, providing a pathway to unify with .

From

Classical ergodic theory examines measure-preserving transformations on probability spaces, where the associated algebra of essentially bounded measurable functions L^\infty(X, \mu) forms a commutative under the trace given by integration against the \mu. This framework captures the through the action of the transformation group on the space, with defined by the triviality of functions up to constants, reflecting unique ergodicity in the commutative case. In the 1970s, this theory extended to the noncommutative setting through the study of group actions on algebras, pioneered by Takesaki, who introduced crossed product constructions to model such dynamics. These crossed products generalize the group-measure-space construction, allowing the analysis of automorphisms induced by group actions on arbitrary algebras rather than just commutative ones. A central in this noncommutative is the , where the fixed-point algebra under the is trivial, consisting only of scalar multiples of the ; this parallels the classical but applies to noncommutative algebras, enabling the study of irreducible dynamics without fixed points. Such actions facilitate the decomposition and of von Neumann algebras via their invariant substructures. The primary motivation from this perspective is the reconstruction of underlying "spaces" from purely algebraic data, particularly in cases of infinite measures where no faithful exists, as exemplified by Connes' of type III factors in the , which uses the flow of weights to recover the modular dynamics and modular invariants. This approach reveals how type III factors, arising from on infinite measure spaces, can be stratified by their associated flows, providing a measure-theoretic analog to geometric reconstruction in noncommutative settings.

Foundational Structures

C*-Algebras

A is a complex algebra A equipped with an involution a \mapsto a^* and a norm \|\cdot\| making A a , such that A is closed under the involution, the norm is submultiplicative (\|ab\| \leq \|a\| \|b\| for all a, b \in A), and satisfies the C*-identity \|a^* a\| = \|a\|^2 for all a \in A. This structure ensures that the involution is continuous and that positive elements (those of the form a^* a) have non-negative spectra, providing a framework for modeling continuous functions on noncommutative spaces. The Gelfand-Naimark theorem establishes that every is isometrically *-isomorphic to a closed *-subalgebra of bounded linear operators on a , linking abstract algebraic properties to concrete operator representations. In the commutative case, the spectrum of a unital C*-algebra A, consisting of nonzero *-homomorphisms from A to \mathbb{C}, is homeomorphic to a compact Hausdorff space X, and the Gelfand transform \hat{A}: A \to C(X) given by \hat{a}(\chi) = \chi(a) is a -isomorphism, recovering the algebra of continuous functions on X. For noncommutative C-algebras, the primitive ideal space \mathrm{Prim}(A), the set of kernels of irreducible *-representations of A equipped with the hull-kernel topology, serves as a noncommutative analogue of the , capturing the topological structure via the space of primitive ideals. This space allows for a hull-kernel correspondence where closed ideals correspond to open subsets, enabling a noncommutative topology. Key examples illustrate the noncommutative nature of these . The C(X) of continuous complex-valued functions on a compact X, with the sup norm \|f\| = \sup_{x \in X} |f(x)| and involution \overline{f(x)}, is a commutative isomorphic via the Gelfand transform to its own . Finite-dimensional examples include the full M_n(\mathbb{C}) over \mathbb{C}, which is a unital with the induced from \mathbb{C}^n and the , representing the "noncommutative points" in finite settings. A prominent infinite-dimensional noncommutative example is the irrational A_\theta, the universal generated by unitaries u and v satisfying vu = e^{2\pi i \theta} uv for irrational \theta \in (0,1), which models the noncommutative and arises from the irrational on . Functional analytic tools underpin the in . Every possesses a contractive approximate unit, a net \{e_\lambda\} such that \|a e_\lambda - a\| \to 0 and \|e_\lambda a - a\| \to 0 for all a \in A, facilitating limits and completions in representations. Closed two-sided ideals I in A are themselves , and the quotient A/I inherits a unique C*-norm, corresponding to "restriction" to the open set \mathrm{Prim}(A) \setminus \mathrm{hull}(I) in the primitive ideal space, thus formalizing open sets and subspaces in noncommutative .

Von Neumann Algebras

Von Neumann algebras, introduced by in the late , are defined as weakly closed *-subalgebras of the bounded operators \mathcal{B}(H) on a H that contain the identity operator. This closure is taken with respect to the weak operator topology, where a net of operators T_\alpha converges weakly to T if \langle T_\alpha \xi, \eta \rangle \to \langle T \xi, \eta \rangle for all \xi, \eta \in H. A fundamental characterization is provided by , which states that for a *-subalgebra \mathcal{M} \subseteq \mathcal{B}(H) containing the identity and closed under adjoints, \mathcal{M} is a if and only if \mathcal{M} = \mathcal{M}'', where \mathcal{M}' = \{ T \in \mathcal{B}(H) \mid T S = S T \ \forall S \in \mathcal{M} \} is the commutant of \mathcal{M} and \mathcal{M}'' is the double commutant. This theorem equates the analytic (topological closure) and algebraic (double commutant) definitions, highlighting the nature of these algebras. Von Neumann algebras are classified into types based on their projection lattices and the existence of traces, a scheme developed by and . Type I algebras are those isomorphic to \mathcal{B}(H) or finite direct sums thereof, subdivided into finite (type I_n for dimension n) and infinite (type I_\infty) cases; their centers consist of diagonal operators with respect to some basis. Type II algebras admit a trace but are infinite-dimensional in a specific sense: type II_1 factors have a finite trace normalizing to 1 on minimal projections, while type II_\infty factors are infinite tensor products of type II_1 with type I_\infty. Type III algebras lack a semifinite trace, characterized by the absence of non-zero finite projections with finite trace; they are further subdivided into type III_\lambda for \lambda \in (0,1] based on the modular spectrum. A factor is a von Neumann algebra with trivial center (just scalar multiples of the identity), and the type classification applies primarily to factors, with general algebras decomposing into direct integrals of factors. Central to von Neumann algebras are projections, which form an orthomodular analogous to measurable sets in classical measure theory, enabling a "noncommutative geometry" of and measure. Two projections p, q \in \mathcal{M} are equivalent if there exists a partial v \in \mathcal{M} with v^* v = p and v v^* = q, defining a function on the . In type II_1 factors, there exists a unique faithful normal semifinite trace \tau: \mathcal{M}^+ \to [0,\infty] such that \tau(1) = 1 and \tau(a b) = \tau(b a) for a, b \in \mathcal{M}, which is finite on the entire positive part and normal (continuous in the ultraweak topology). This trace extends the classical to noncommutative settings, with \tau(p) giving the "" of projection p. The Tomita-Takesaki theory provides a dynamical framework for via states, constructing a one-parameter group of automorphisms from a faithful \phi. For a \mathcal{M} in standard form on a with cyclic and separating vector \Omega implementing \phi(a) = \langle a \Omega, \Omega \rangle, the Tomita operator S: \mathcal{M} \Omega \to \mathcal{M} \Omega defined by S a \Omega = a^* \Omega has polar decomposition S = J \Delta^{1/2}, where J is an antiunitary and \Delta is the modular operator. The modular automorphism group \{\sigma_t^\phi\}_{t \in \mathbb{R}} is then given by \sigma_t^\phi(a) = \Delta^{i t} a \Delta^{-i t} for a \in \mathcal{M}, preserving \phi and satisfying the condition at inverse temperature 1. This theory links to , where states on C*-dynamical systems describe thermal equilibrium, with the modular flow recovering the for type III factors in . arise as weak closures of concrete representations on , bridging topological and measurable aspects of noncommutative geometry.

Noncommutative Geometric Constructions

Noncommutative Differentiable Manifolds

Noncommutative differentiable manifolds generalize the notion of structures on classical manifolds to noncommutative algebras, replacing bundles with derivations and calculi that satisfy a noncommutative . In this framework, the algebra A plays the role of functions on the "space," and derivations \delta: A \to E into a bimodule E encode changes, obeying \delta(ab) = \delta(a)b + a\delta(b) for a, b \in A. This allows the construction of higher forms via iterated applications, extending to noncommutative settings. Smooth algebras form the basis for these structures, typically defined as Fréchet or locally algebras over \mathbb{C}, equipped with a compatible with algebraic operations. For instance, the algebra C^\infty(\Sigma) of functions on a manifold \Sigma is a prototypical smooth algebra, but noncommutative examples include dense subalgebras of C^*-, such as the smooth noncommutative torus A_\theta^\infty = \{ \sum a_{nm} U_1^n U_2^m \mid a_{nm} \in \mathcal{S}(\mathbb{Z}^2) \}, where \mathcal{S}(\mathbb{Z}^2) denotes rapidly decreasing sequences and U_1, U_2 are unitary generators satisfying U_1 U_2 = e^{2\pi i \theta} U_2 U_1. These algebras admit continuous derivations, enabling a differential calculus that mimics smooth geometry while accommodating noncommutativity. The universal differential envelope \Omega^1(A) provides a canonical way to associate a module of 1-forms to any unital A, realized as the A-bimodule \ker(m: A \otimes_k A \to A), with bimodule structure a \cdot (u \otimes v) \cdot b = (a u) \otimes (v b), and the universal derivation d: A \to \Omega^1(A) given by d(a) = a \otimes 1 - 1 \otimes a. This construction is universal among bimodules E admitting derivations from A, meaning any derivation \delta: A \to E factors uniquely through \Omega^1(A). Higher forms \Omega^n(A) are obtained by tensoring, forming a graded differential with d^2 = 0, which supports noncommutative analogs of integration and homology. Connections on these manifolds are defined as bimodule maps \nabla: E \to \Omega^1(A, E) from a right A-module E to the space of 1-forms with values in E, satisfying a Leibniz rule \nabla(\xi a) = \nabla(\xi) a + \xi \otimes da for \xi \in E, a \in A. The curvature of such a connection is the A-bimodule map \Theta = \nabla^2: E \to \Omega^1(A, E), extended bilinearly over \Omega^1(A) to define F_\nabla(\xi, \eta) \in \End_A(E) \otimes_A \Omega^2(A), where \Theta^2 = 0 in flat cases. For projective modules over smooth algebras like A_\theta^\infty, connections can be chosen with constant , such as F_\nabla = -2\pi i \varepsilon \otimes (e_1 \wedge e_2) on the basic line bundle, facilitating Yang-Mills theory in noncommutative settings. A representative example is the Weyl algebra W_n, the unital associative algebra over \mathbb{R} generated by position operators q_1, \dots, q_n and momentum operators p_1, \dots, p_n satisfying the canonical commutation relations [q_i, p_j] = i \hbar \delta_{ij} and [q_i, q_j] = [p_i, p_j] = 0, which models the noncommutative phase space \mathbb{R}^{2n}. This algebra admits a rich differential structure via derivations like the Euler derivation \delta(a) = \sum (q_i \partial_{q_i} a + p_i \partial_{p_i} a), and its universal differential envelope captures the noncommutative symplectic geometry. Another key example is Connes' tangent groupoid G_M^t for a smooth manifold M, constructed as the deformation (t, x, y, v) \in [0,1] \times M \times M \times T_x M with groupoid multiplication smoothing the tangent bundle TM (at t=0) into the pair groupoid M \times M (at t=1), used to resolve singularities in differential operators. Noncommutative coordinates in these manifolds are elements of the A that behave like deformed position functions, satisfying a twisted Leibniz rule under derivations in quantized settings, such as d(ab) = d(a) \star b + a \star d(b) where \star is a star product deforming the pointwise multiplication. For the noncommutative , coordinates are the generators U_1, U_2, with derivations \partial_1(U_1) = 2\pi i U_1, \partial_1(U_2) = 0, and similarly for \partial_2, ensuring compatibility with the . This framework briefly references algebraic varieties by viewing smooth as completions of coordinate rings, though the focus remains on analytic aspects.

Noncommutative Affine and Projective Schemes

Noncommutative affine schemes generalize the classical notion of affine schemes from commutative algebraic geometry to the setting of noncommutative rings. In this framework, an affine scheme is associated to a noncommutative algebra A, where the "points" of the scheme are interpreted through the primitive spectrum, consisting of primitive ideals, or equivalently, via irreducible representations of A. This construction allows for a categorical description where the geometry is captured by the category of quasicoherent modules over A, extending the functorial approach of commutative schemes. Key examples of such noncommutative include extensions and skew polynomial , which provide noncommutative analogues of . An extension of a R is formed by adjoining an indeterminate x with a \sigma- \delta, yielding R[x; \sigma, \delta], where multiplication satisfies x r = \sigma(r) x + \delta(r) for r \in R. These satisfy the Ore condition, enabling localizations and fraction fields, and are used to model affine with twisted actions, such as in quantum coordinate . Skew polynomial , a special case where \delta = 0, further exemplify domains whose maximal right ideals correspond to points in the associated affine . Projective modules over noncommutative rings play the role of vector bundles in this . For noncommutative domains like the quantum plane A = k\langle x, y \rangle / (yx - q xy), where q \neq 0 is a scalar, finitely generated projective modules admit a well-defined , defined locally via or generalized to the global function. Stability of these modules refers to the property that projectives of constant n become isomorphic after stabilization by modules, facilitating the study of noncommutative vector bundles. This and stability enable classifications analogous to classical , with the quantum plane serving as a prototypical example where all projectives are of equal to their . Van den Bergh duality provides a bridge between projective and affine noncommutative varieties, establishing a Poincaré-type duality relating and for s. For a noncommutative A, this duality asserts that \mathrm{HH}_i(A) \cong \mathrm{HH}^{d-i}(A, A^\circ)^\vee for some d, where A^\circ is the opposite , and it extends to relate the derived categories of coherent sheaves on projective schemes (via graded quotients) to those on affine covers. This framework unifies affine and projective constructions, allowing global properties of noncommutative varieties to be derived from local affine data. Prominent examples include s, whose structures serve as coordinate rings for noncommutative affine s, encoding symmetries in deformed spaces. For instance, the coordinate algebra O_q(SL_2) defines an affine with points given by representations. In the projective setting, Manin's noncommutative projective space \mathbb{P}^n_q is constructed as the Proj of the q-skew in n+1 variables, where relations are x_i x_j = q_{ij} x_j x_i for i < j, modeling twisted with modules corresponding to quantum line bundles.

Invariants and Topology

K-Theory and Cyclic Cohomology

In noncommutative geometry, algebraic K-theory provides a key topological invariant for noncommutative algebras, generalizing the study of vector bundles on spaces to modules over algebras. For an associative unital algebra A, the zeroth algebraic K-group K_0(A) is defined as the Grothendieck group of the abelian monoid of isomorphism classes of finitely generated projective left A-modules under direct sum, or equivalently, as the abelian group generated by equivalence classes of idempotents in the matrix algebras M_n(A) for n \geq 1, subject to the relation + = [e \oplus f] where idempotents e, f are stably equivalent if there exists g such that e \oplus g \sim f \oplus g. Higher algebraic K-groups K_n(A) for n \geq 1 are defined using Quillen's + construction on the classifying space BGL(A)^+ of the infinite general linear group GL(A) = \varinjlim GL_n(A), yielding K_n(A) = \pi_n(BGL(A)^+), which captures homotopy-theoretic information about the stable range of matrices over A. This framework extends classical Grothendieck's K-theory of schemes to noncommutative settings, enabling the classification of projective modules and their stable isomorphism classes. For C*-algebras, which model noncommutative topological spaces, adapts the algebraic version to the continuous setting, emphasizing stable homotopy classes of projections. The groups K_0(A) and K_1(A) for a unital A are defined as the Grothendieck group of Murray-von Neumann equivalence classes of projections in M_\infty(A) = \varinjlim M_n(A) and the quotient GL_\infty(A)/[GL_\infty(A), GL_\infty(A)], respectively, with higher groups vanishing in basic cases but revived by periodicity. A hallmark is Bott periodicity, which asserts that K_{n+2}(A) \cong K_n(A) for all n \geq 0, establishing a \mathbb{Z}/2\mathbb{Z}-periodic structure analogous to the classical Bott theorem for complex vector bundles on spheres; this isomorphism arises from suspension functors and the exact sequence of the ideal of compact operators. In noncommutative geometry, these K-groups classify stable isomorphism classes of projections, providing invariants like the range of the trace on projections in AF-algebras. Cyclic cohomology introduces a complementary to , capturing periodic structures in the Hochschild cohomology of algebras. For an A, cyclic cohomology HC^n(A) is the cohomology of the cyclic complex CC^*(A), obtained by imposing cyclic invariance b' + B = 0 on the Hochschild cochain complex C^*(A, A^*) with s b (Hochschild ) and B (cyclic derived from the Connes-Tsygan of cyclic objects). Connes' long relates it to Hochschild cohomology HH^n(A): \cdots \to HC^{n}(A) \xrightarrow{I} HH^{n}(A) \xrightarrow{B} HC^{n-1}(A) \xrightarrow{S} HC^{n+1}(A) \to \cdots, where I is the inclusion, B the connecting homomorphism, and S the periodicity operator S(\phi)(a_0, \dots, a_n) = \sum (-1)^k \phi(a_k, \dots, a_0, a_k), ensuring HC^n(A) \oplus HC^{n+1}(A) \cong HC^{n+2}(A) in even-odd periodicity. This sequence, arising from the short exact sequence of complexes $0 \to \mathrm{Im}(b') \to C^*(A)/\mathrm{Im}(b') \to CC^*(A) \to 0, computes cyclic cohomology from Hochschild data and underlies tools like the Connes-Gysin sequence for filtered algebras. The Chern-Connes character establishes a deep pairing between these invariants, generalizing the classical Chern character from to . Defined as a natural transformation \mathrm{ch}: K_*(A) \to HC_{\mathrm{even}}^*(A), it maps a K-theory class \in K_0(A) represented by a projection p to the cyclic cocycle \mathrm{ch}(p) = \sum_{k=0}^\infty \frac{(-1)^k}{(2k)!} \mathrm{Tr}(p (dp)^{2k}) in the even periodic cyclic cohomology, where d = [D, \cdot] for a generalized differential; in the algebraic case, it uses idempotent traces. This character is a ring homomorphism compatible with Bott periodicity and induces the index pairing \langle \mathrm{ch}([E]), [\tau] \rangle = \mathrm{Ind}(D_E) for traces \tau \in HC^0(A) and Dirac-type operators D_E, recovering the Atiyah-Singer index theorem in noncommutative settings. The pairing \langle \alpha, \beta \rangle = \int \alpha \cup \beta for \alpha \in HC^*(A), \beta \in K_*(A) via the character, yields numerical invariants like local formulas for the Euler characteristic in cyclic terms.

Spectral Triples and Dirac Operators

A spectral triple is the central object in noncommutative geometry that encodes both metric and differential structure on a noncommutative space in an operator-theoretic framework. Formally, it consists of a triple (\mathcal{A}, \mathcal{H}, D), where \mathcal{A} is a unital involutive algebra acting faithfully by adjointable operators on the Hilbert space \mathcal{H}, and D is an unbounded self-adjoint operator on \mathcal{H} serving as a Dirac operator, with the key condition that the commutator [D, a] is a bounded operator on \mathcal{H} for all a \in \mathcal{A}. This condition ensures that the commutators behave like one-forms in the noncommutative setting, allowing the construction of a differential calculus. Additionally, D is typically required to have compact resolvent (D - \lambda I)^{-1} for \lambda \notin \mathbb{R}, leading to a discrete spectrum of eigenvalues, which mimics the spectral properties of elliptic operators on manifolds. The Dirac operator D induces a natural metric on the state space of \mathcal{A} through Connes' distance formula: for pure states \phi, \psi corresponding to points in the classical limit, the distance is given by d(\phi, \psi) = \sup \bigl\{ |\phi(a) - \psi(a)| : a \in \mathcal{A}, \ \| [D, a] \| \leq 1 \bigr\}, which recovers the geodesic distance on the underlying space when \mathcal{A} is commutative. This metric provides a way to measure distances in noncommutative geometries, generalizing the infimum of lengths of curves connecting points, and it is Lipschitz continuous with respect to the operator norm. The formula highlights how the spectrum of D controls the geometry, as larger eigenvalues correspond to finer scales. In the commutative case, where \mathcal{A} = C^\infty(M) for a compact Riemannian spin manifold M, the Hilbert space \mathcal{H} = L^2(M, S) consists of square-integrable spinor sections, and D is the classical associated to the spin structure. This spectral triple precisely recovers the of M, with the distance formula yielding the geodesic distance and the commutators [D, f] corresponding to Clifford multiplication by the differential df. The eigenvalues of D encode the metric via Weyl's law, linking the spectral data to the volume and curvature. For noncompact or discrete spaces, such as those arising from s, spectral triples incorporate conditions of bounded to ensure finite propagation speed in the associated . These conditions require that the resolvent of D lies in suitable Schatten ideals and that the exhibits uniform bounds on local features, like injectivity radius and , allowing the \operatorname{Tr}(e^{-D^2}) to be finite and mimicking estimates on manifolds of bounded . Examples include the spectral triple on the reduced group C*- of a \Gamma, where D is constructed to capture the coarse of \Gamma.

Examples and Applications

Specific Noncommutative Spaces

One prominent example of a noncommutative space is the noncommutative torus A_\theta, which serves as a deformation of the classical two-dimensional torus. This space is defined as the universal unital C^*-algebra generated by two unitary elements u and v satisfying the commutation relation uv = e^{2\pi i \theta} vu, where \theta \in \mathbb{R} is a deformation parameter, often taken to be irrational to ensure simplicity. The algebra A_\theta captures the noncommutative analogue of periodic functions on the torus, with its structure reflecting rotational symmetry deformed by \theta. A spectral triple for A_\theta can be constructed using the representation on the Hilbert space associated with the discrete Heisenberg group, where the Dirac operator incorporates the twisted group action to encode a metric and differential structure. The quantum provides another concrete realization, arising as a of the SU_q(2) by the subgroup SU_q(1), where q is a deformation parameter with |q| \neq 1. This noncommutative two- S^2_q is realized within the algebra of functions on SU_q(2), using a that preserves covariance under the action. The coordinate algebra is generated by elements satisfying quantum analogues of the classical relations, such as x_1^2 + x_2^2 + x_3^2 = 1 deformed by q-commutators. A differential calculus on S^2_q is equipped with braided symmetry, where the exterior derivative and wedge product respect a braiding derived from the U_q(\mathfrak{sl}(2))-module structure, enabling computations of noncommutative integrals and volumes. Finite noncommutative geometries model spaces through finite-dimensional , particularly M_n(\mathbb{C}), which act as the algebra of functions on a "finite point" or a set of n indistinguishable points. In this framework, M_n(\mathbb{C}) acts on the finite-dimensional \mathbb{C}^n, and the geometry is specified by a D, a with finite spectrum that encodes and pairings via the spectral triple (M_n(\mathbb{C}), \mathbb{C}^n, D). For instance, in the two-point space, the algebra \mathbb{C} \oplus \mathbb{C} pairs with a yielding an on the order of $10^{-16} cm, illustrating how such structures capture electroweak scales in models. The resolvent of D is compact by finiteness, and the [D, a] for a \in M_n(\mathbb{C}) remains bounded, ensuring the triple's validity. Noncommutative analogues of the three-sphere extend these ideas to higher dimensions with relations incorporating deformation parameters. These spaces, studied as moduli spaces of classes of algebras, feature central forms that classify their structure up to . For example, the noncommutative S^3 is parameterized by pairs of points in a symmetric space of unitary matrices, with the algebra generated by coordinates satisfying deformed relations like those in SU_q(2). This construction yields a family of three-dimensional noncommutative manifolds whose real and complex moduli spaces are explicitly described, providing insights into their topological and analytic properties without continuous symmetries.

Applications in Physics and Index Theory

Noncommutative geometry provides a framework for generalizing the to noncommutative manifolds through spectral triples ( \mathcal{A}, \mathcal{H}, [D](/page/D*) ), where D is a Dirac-like operator. The local index formula arises from the analytic index computed via cyclic cohomology, pairing the Chern character in with residues of the zeta function \zeta_D(s) = \Tr |D|^{-s}. This extends the classical theorem by incorporating noncommutative invariants, such as the pairing \langle [\text{ch}(\cdot)], \tau \rangle, where \tau is a cyclic cocycle derived from the heat kernel expansion. A key tool is the spectral action \Tr f(D / \Lambda), with f a smooth cutoff function, often approximated by the heat trace \Tr e^{-D^2 / \Lambda^2}. The asymptotic expansion as \Lambda \to \infty yields a series in even powers of \Lambda, with leading terms reproducing the Einstein-Hilbert action for coupled to Yang-Mills-Higgs fields. Specifically, the expansion includes the term a_0 \Lambda^4 \int \sqrt{g} \, d^4x and the Higgs potential from quadratic fluctuations, unifying and in an Einstein-Higgs system. This local formula generalizes the Atiyah-Singer theorem by providing a trace anomaly that encodes index densities on noncommutative spaces. In , emerges from an almost commutative spectral triple, constructed as a tensor product of the classical spectral triple and a finite-dimensional noncommutative \mathcal{F} of KO-dimension 6. The finite part \mathcal{F} is the algebra of matrices encoding generations and representations under the gauge group U(1) \times SU(2) \times SU(3), with a real structure ensuring the correct content, including right-handed neutrinos. The full D = D_M \otimes 1 + \gamma^5 \otimes D_F reconstructs the via the spectral action and a fermionic term, yielding the bosonic sector with gauge fields and the Higgs doublet. The Higgs field arises naturally as an inner fluctuation of the , given by D_\phi = D + \phi + J \phi^* J^{-1}, where \phi = h(x) \otimes 1_F + \dots incorporates the finite elements, and J is the real structure. This fluctuation generates the Higgs potential V(h) = m^2 |h|^2 + \lambda |h|^4 from the spectral expansion. Extensions to include neutrino mixing adjust \mathcal{F} to incorporate Majorana masses. Applications to include the , where noncommutative geometry models the lowest Landau level as a noncommutative . The integer Hall conductivity is quantized as a noncommutative Chern number, given by the pairing \sigma_H = \frac{e^2}{h} \langle , \text{ch}_2(\tau) \rangle in , with $$ the idempotent projector and \tau a cyclic cocycle from the trace on the irrational rotation . For the , twisted on hyperbolic geometries captures Laughlin states, expressing filling factors \nu = p/q as ratios of Euler characteristics. Recent extensions in the 2020s incorporate deformation quantization for inhomogeneous , revealing noncommutative structures in the Girvin-MacDonald-Platzman of operators. This explains fractional excitations via and Fedosov star-products, linking to states and statistics in higher-genus surfaces. In gravity and physics, the action on quantized spacetimes, modeled by deformed , yields effective actions for noncommutative metrics. For , this leads to regularized horizons via smeared mass distributions, avoiding singularities while preserving asymptotic flatness. The action's expansion includes higher-derivative terms like R^2, stabilizing quantum corrections. Recent 2020s research connects this to asymptotic safety, where the action's running couplings approach a UV fixed point, ensuring renormalizability; generalized principles from noncommutativity align with Reuter fixed-point trajectories, predicting modified evaporation and shadows observable in .

Advanced Connections

Connes' Notion of Connection

In noncommutative geometry, defines a connection on a E over an A as a \nabla: E \to E \otimes_A \Omega^1(A), where \Omega^1(A) denotes the space of one-forms, satisfying the Leibniz rule \nabla(\xi a) = (\nabla \xi) a + \xi \otimes da for \xi \in E and a \in A. This generalization extends classical to noncommutative settings, allowing connections to act on modules that replace vector bundles. The of such a connection is given by \Omega = \nabla^2, which measures the deviation from flatness and takes values in endomorphisms of E \otimes_A \Omega^1(A). Within the framework of spectral triples, inner fluctuations of the connection arise from commutators [D, \pi(f)] for f \in A, where D is the and \pi is the representation of A on the . These fluctuations generate gauge potentials, enabling the construction of perturbed Dirac operators D + [D, \pi(f)] that preserve the spectral triple structure while incorporating transformations. Metric connections in this context are those compatible with the , ensuring a notion of length and on noncommutative spaces. The associated Yang-Mills is formulated as \int \operatorname{Tr} |\Omega|^2, where the is a noncommutative often involving the Dixmier on pseudodifferential operators of -d in d. This quantifies the of the curvature and plays a central role in deriving for gauge fields. Connes' approach further unifies theories with through bimodule derivations, where the one-forms \Omega^1_D are generated by [D, a] for a \in A, providing a bimodule structure that links noncommutative potentials to gravitational effects. This unification manifests in models where the algebra decomposes into finite-dimensional components, such as A = \mathbb{C} \oplus \mathbb{H} \oplus M_3(\mathbb{C}), yielding the group U(1) \times SU(2) \times SU(3) alongside the group for . Noncommutative geometry establishes deep connections with quantum groups, which are typically realized as that deform classical and their representations. A foundational example is the quantized universal enveloping algebra U_q(\mathfrak{g}), where \mathfrak{g} is a and q is a deformation parameter; this structure was introduced independently by Drinfeld and as a q-deformation preserving the axioms while introducing noncommutativity in the relations among generators. Similarly, the quantum coordinate algebra O(SU_q(2)) deforms the algebra of functions on the SU(2), providing a noncommutative analogue of the classical that serves as a building block for more complex quantum spaces in noncommutative geometry. These quantum groups enable the description of symmetries on noncommutative spaces, where the in the encodes how group actions distribute over tensor products of representations. Bicrossed products further link quantum groups to noncommutative spaces by constructing new s from matched pairs of algebras and coalgebras, yielding examples like the quantum double or Poincaré quantum group that model deformed . Developed by Majid, this construction produces noncommutative geometries such as quantum homogeneous spaces, where the bicrossed product A \bowtie H combines a A with a H via mutual actions and coactions, resulting in a noncocommutative yet quasitriangular structure. In noncommutative geometry, bicrossed products facilitate the quantization of group manifolds, allowing the recovery of classical limits as the deformation approaches 1, and they underpin applications to quantum field theories on curved noncommutative backgrounds. Deformation quantization bridges noncommutative geometry with geometry by associating to any a family of noncommutative algebras via star products. Kontsevich's formality theorem establishes that every finite-dimensional admits a deformation quantization, where the star product f \star_\hbar g = f g + \sum_{n \geq 1} \hbar^n B_n(f,g) (with B_n bidifferential operators) deforms the pointwise multiplication, and the \hbar \to 0 recovers the commutative . This formality, proven through a combinatorial graph expansion and quasi-isomorphism between Hochschild cochains and Gerstenhaber brackets, classifies all star products up to equivalence and extends Fedosov's construction from to general settings. Drinfeld twists provide a mechanism to deform commutative algebras into noncommutative ones using 2-cocycles on a , preserving the module structure while introducing noncommutativity via a twist element \mathcal{F} \in H \otimes H satisfying the cocycle condition. In noncommutative geometry, applying a Drinfeld twist to the universal enveloping algebra U(\mathfrak{g}) of a Lie algebra \mathfrak{g} yields a quasi-Hopf algebra whose representations deform the classical differential geometry, such as twisting the commutative coordinate algebra of a manifold to a noncommutative version. This approach, originating from Drinfeld's work on quasi-triangular structures, allows systematic quantization of Poisson-Lie groups and has been extended to non-equivariant connections on bimodules, enabling deformations of principal bundles and gauge theories. Recent advances in the 2020s have integrated into noncommutative symplectic geometry, particularly in exploring mirror symmetry for deformed Calabi-Yau spaces. For instance, quantum toric varieties, constructed as noncommutative deformations using quantum group actions, provide symplectic structures that dualize under mirror symmetry, linking algebraic and geometric quantizations via symmetries. These developments extend Kontsevich's formality to noncommutative settings with quantum group twists, facilitating in derived categories of coherent sheaves over quantum spaces.

References

  1. [1]
    [PDF] Noncommutative Geometry Alain Connes
    ... basic idea of algebraic geometry. The purpose of this book is to extend this correspondence to the noncommutative case in the framework of real analysis.
  2. [2]
    [PDF] A short introduction to noncommutative geometry
    Nov 23, 2004 · Definition: A 'noncommutative' spectral triple is a system (H, ˆA, D) consisting of the following objects: • a Hilbert space H,. • a ...
  3. [3]
    [1204.0328] Particle Physics from Almost Commutative Spacetimes
    Apr 2, 2012 · Abstract:Our aim in this review article is to present the applications of Connes' noncommutative geometry to elementary particle physics.<|control11|><|separator|>
  4. [4]
  5. [5]
    [2207.10901] Noncommutativity and Physics: A non-technical review
    Jul 22, 2022 · We give an overview of the applications of noncommutative geometry to physics. Our focus is entirely on the conceptual ideas, rather than on the underlying ...
  6. [6]
    [PDF] Notes on the Gelfand-Neumark theorem
    Jun 5, 2019 · In 1943, a paper [G-N], written by I. M. Gelfand and M. Neumark, "On the imbedding of normed rings into the ring of operators in Hilbert ...Missing: original | Show results with:original
  7. [7]
    [PDF] Notes on Noncommutative Geometry - Mathematics
    Other events of great importance to noncommutative geometry were the definition of C*-algebras and the first results in this area by Gelfand and Naimark in 1943 ...
  8. [8]
    Approximation in, and Representation of, Certain Banach Algebras
    By RICHARD ARENS. 1. Introduction. The primary purpose of this paper is to extend to non-commutative Banach algebras the functional representation theorems.Missing: program | Show results with:program
  9. [9]
    [PDF] The flashes of insight never came for free
    Dec 4, 2010 · Connes's early work, for which he was award- ed the Fields Medal in 1982, was concerned with the classification of von Neumann alge- bras. This ...
  10. [10]
    The von Neumann algebra of a foliation - SpringerLink
    Every smooth foliation of a manifold V gives rise very naturally to a von Neumann algebra . The weights on M correspond exactly to...
  11. [11]
    Quantum Groups and Noncommutative Geometry - Yuri I. Manin
    Oct 11, 2018 · This textbook presents the second edition of Manin's celebrated 1988 Montreal lectures, which influenced a new generation of researchers in ...Missing: projective | Show results with:projective
  12. [12]
    [q-alg/9709040] Deformation quantization of Poisson manifolds, I
    Sep 29, 1997 · Authors:Maxim Kontsevich. View a PDF of the paper titled Deformation quantization of Poisson manifolds, I, by Maxim Kontsevich. View PDF.
  13. [13]
    Noncommutative homological mirror symmetry of elliptic curves - arXiv
    Apr 20, 2020 · One is the Polishchuk-Zaslow's mirror symmetry functor of elliptic curves, and the other is a localized mirror functor from the Fukaya ...Missing: 2020s | Show results with:2020s
  14. [14]
    [1403.4805] A Course on Noncommutative Geometry in String Theory
    Mar 19, 2014 · In this pedagogical mini course the basics of the derivation of the noncommutative structures appearing in string theory are reviewed.Missing: CFT 2010s
  15. [15]
    [PDF] PARTICLE MODELS AND NONCOMMUTATIVE GEOMETRY - IHES
    We write three particle models in terms of noncommutative gauge theory: the Glashow-. Weinberg-Salam model, the Peccei-Quinn model and the standard model. I.
  16. [16]
    Noncommutative Geometry and the Thermodynamic Fate of Black ...
    Sep 27, 2025 · We show that the conventional first law of black hole thermodynamics is violated since the entropy deviates from the Bekenstein-Hawking form.Missing: spectral action 2020s
  17. [17]
    [PDF] von Neumann algebras and Ergodic Theory
    Apr 10, 2008 · The prototypical example of an abelian von Neumann algebra is given by the algebra A = L∞(X,B,µ) of essentially bounded measurable functions ...Missing: classical | Show results with:classical
  18. [18]
    [PDF] lectures on ergodic theory of group actions (a von neumann algebra ...
    Denote Aut(L∞X, τµ) the group of ∗-automorphisms of the von Neumann algebra L∞X that preserve the functional τµ = / ·dµ, and identify Aut(X, µ) and Aut(L∞X, τµ ...
  19. [19]
    DUALITY IN CROSSED PRODUCTS AND VON NEUMANN ...
    In this paper, we announce further results succeeding to the previous papers [2] and [3]. Suppose Ji is a von Neumann algebra and G an.
  20. [20]
    [PDF] arXiv:1404.5594v1 [math.OA] 22 Apr 2014
    Apr 22, 2014 · The action α is called ergodic if the fixed point algebra is trivial, that is,. Nα := { x ∈ N : α(x) = 1 ⊗ x } = C1 . 2. Convolution ...
  21. [21]
    Une classification des facteurs de type ${\rm III} - EuDML
    Connes, Alain. "Une classification des facteurs de type ${\rm III}$." Annales scientifiques de l'École Normale Supérieure 6.2 (1973): 133-252. <http://eudml.org ...Missing: reconstructing | Show results with:reconstructing<|control11|><|separator|>
  22. [22]
    [PDF] Factors of Type III - Numdam
    We then détermine this flow for a factor given by the so-called group measure space construction. ... Connes, Une classification de facteurs de type III, Ann. Sci ...
  23. [23]
    On the imbedding of normed rings into the ring of operators ... - EuDML
    This paper, by Гельфанд and Наймарк, discusses the imbedding of normed rings into the ring of operators in Hilbert space, published in Matematiceskij sbornik ...Missing: PDF | Show results with:PDF
  24. [24]
    [PDF] Operator Algebras Theory of C - try
    Feb 8, 2017 · This volume attempts to give a comprehensive discussion of the theory of opera- tor algebras (C*-algebras and von Neumann algebras.) ...
  25. [25]
    C*-Algebras and Operator Theory - ScienceDirect.com
    This chapter presents a study of C*-algebras and of operators on Hilbert spaces. Hilbert spaces are very well-behaved compared with general Banach spaces, and ...
  26. [26]
    The Primitive Ideal Space of a C*-Algebra
    Nov 20, 2018 · The commutative Gelfand-Naimark Theorem says that any commutative C*-algebra A is isomorphic to the ring C 0(M, C) of all continuous ...
  27. [27]
    [PDF] Imbedding the irrational rotation C
    Let 0 (0, 1) be an irrational number and A., the corresponding irrational rotation C*-algebra (i.e. the cross-product of the continuous functions on the unit.
  28. [28]
    [PDF] John von Neumann and the Theory of Operator Algebras *
    Von Neumann's double commutant theorem asserts that a family of operators is a von Neumann algebra if and only if it contains the adjoint of its elements and ...
  29. [29]
    ON RINGS OF OPERATORS. II* - American Mathematical Society
    On rings of operators, Annals of Mathematics, (2), vol. 37 (1936), pp. 116-229. It contains the solution of certain problems which were left open there. We.
  30. [30]
    Tomita's Theory of Modular Hilbert Algebras and its Applications
    Free delivery 14-day returnsTable of contents (15 chapters) · Front Matter. Pages I-II. Download chapter PDF · Introduction. M. Takesaki · Preliminaries. M. Takesaki · Modular Hilbert algebras.
  31. [31]
    [PDF] The Role of Type III Factors in Quantum Field Theory - arXiv
    Dec 4, 2004 · The mathematical problem addressed by Murray and von Neumann was to classify all possibilities for such algebras. This problem is motivated by ...
  32. [32]
    [PDF] Non-commutative differential geometry - Numdam
    This is the introduction to a series of papers in which we shall extend the calculus of differential forms and the de Rham homology of currents beyond their ...
  33. [33]
    [PDF] arXiv:math/0002177v1 [math.RA] 22 Feb 2000
    SMOOTH ALGEBRAS IN NONCOMMUTATIVE GEOMETRY. 5 to commutative algebras. One checks that if A = SV/I is any presentation of A as a quotient of a symmetric ...<|control11|><|separator|>
  34. [34]
    [PDF] Noncommutative smooth spaces - arXiv
    Before going further on we would like to notice that TA is isomorphic as an abstract Z- graded algebra to the so-called differential envelope ΩA, the universal ...
  35. [35]
    [PDF] Noncommutative Differential Geometry of Generalized Weyl Algebras
    Jun 23, 2016 · The resulting commutative algebra is the coordinate algebra of a parallelisable manifold (the product of the circle with the plane or the ...
  36. [36]
    [PDF] Twisting all the way - arXiv
    Oct 31, 2007 · u(h⋆g) = L⋆ u(h)⋆g + ¯Rα(h)⋆L⋆ ¯Rα(u)(g) . This deformed Leibniz rule is intuitive: in the second addend we have exchanged the order of u and h ...
  37. [37]
    [PDF] The K-book an introduction to Algebraic K-theory
    Aug 29, 2013 · Algebraic K-theory has two components: the classical theory which centers around the Grothendieck group K0 of a category and uses explicit ...
  38. [38]
    [PDF] arXiv:1803.01551v2 [math.OA] 4 May 2018
    May 4, 2018 · of bounded geometry by similar methods [55, Section 5]. It is hoped ... Now we prove that the triple (3.10) is a spectral triple satisfying Hypoth ...
  39. [39]
    [PDF] Lectures on Noncommutative Geometry - arXiv
    Apr 12, 2007 · ... and let λjk = e2πiΘjk . The noncommutative torus AΘ is the universal unital C∗-algebra generated by unitaries U1,U2,··· ,Un with. UkUj ...
  40. [40]
    None
    Summary of each segment:
  41. [41]
    None
    ### Summary of Quantum Sphere as SU_q(2) Quotient and Differential Calculus with Braided Symmetry
  42. [42]
    [PDF] DIFFERENTIAL CALCULI ON THE QUANTUM GROUP SUq ... - arXiv
    A variety of three-dimensional left-covariant differential calculi on the quantum group SUq(2) is considered using an approach based on global U(1)-.
  43. [43]
    Moduli space and structure of noncommutative 3-spheres - arXiv
    Aug 28, 2003 · The moduli space of noncommutative 3-spheres is identified with equivalence classes of pairs of points in a symmetric spaceof unitary ...Missing: Mulcahy | Show results with:Mulcahy
  44. [44]
    The local index formula in noncommutative geometry
    In noncommutative geometry a geometric space is described from a spectral vantage point, as a tripleA, H, D consisting of a *-algebraA represented in a Hil.Missing: action | Show results with:action
  45. [45]
    [hep-th/9606001] The Spectral Action Principle - arXiv
    Jun 3, 1996 · Title:The Spectral Action Principle. Authors:Ali H. Chamseddine, Alain Connes (ETH and I.H.E.S). View a PDF of the paper titled The Spectral ...
  46. [46]
  47. [47]
    [0706.3688] Why the Standard Model - arXiv
    Jun 25, 2007 · The Standard Model is based on the gauge invariance principle with gauge group U(1)xSU(2)xSU(3) and suitable representations for fermions and bosons.
  48. [48]
    [hep-th/0605011] Inner fluctuations of the spectral action - arXiv
    May 1, 2006 · We prove in the general framework of noncommutative geometry that the inner fluctuations of the spectral action can be computed as residues and ...
  49. [49]
    [cond-mat/0502356] Towards the fractional quantum Hall effect - arXiv
    Feb 15, 2005 · Abstract: In this paper we give a survey of some models of the integer and fractional quantum Hall effect based on noncommutative geometry.Missing: Chern K- 2020s
  50. [50]
    Noncommutative Geometry and Deformation Quantization in ... - arXiv
    Apr 8, 2023 · Here, we discuss the underlying noncommutative geometry of quantum Hall fluids in which the magnetic fields are spatially inhomogenoeus.Missing: developments 2020s
  51. [51]
  52. [52]
    [1909.02434] Black hole entropy from trace dynamics and non ...
    Sep 5, 2019 · In our theory, which we call non-commutative matter gravity, a black hole arises from the spontaneous localisation of an entangled state of a ...
  53. [53]
    [hep-th/0006167] Quantum Groups and Noncommutative Geometry
    Jun 21, 2000 · Quantum Groups and Noncommutative Geometry. Authors:S. Majid. View a PDF of the paper titled Quantum Groups and Noncommutative Geometry, by S.Missing: bicrossed | Show results with:bicrossed
  54. [54]
    Extensions of locally compact quantum groups and the bicrossed ...
    In the framework of locally compact quantum groups, we study cocycle actions. We develop the cocycle bicrossed product construction, starting from a matched ...
  55. [55]
    [PDF] Deformation quantization of Poisson manifolds - IHES
    Here is the final version of the e-print Deformation quantization of Poisson manifolds I [33] posted on the web archive as q-alg/9709040.
  56. [56]
    [1210.0241] Noncommutative connections on bimodules and ... - arXiv
    Sep 30, 2012 · We extend the Drinfeld twist deformation theory of modules and algebras to morphisms and connections that are not necessarily H-equivariant.
  57. [57]
    Quantum (non-commutative) toric geometry: Foundations
    Nov 19, 2021 · In this paper, we will introduce Quantum Toric Varieties which are (non-commutative) generalizations of ordinary toric varieties where all ...