Fact-checked by Grok 2 weeks ago

Projective Hilbert space

In mathematics and physics, the projective Hilbert space associated with a complex Hilbert space H is the set of equivalence classes of non-zero vectors in H, where two vectors \psi and \phi are considered equivalent if \psi = c \phi for some non-zero complex scalar c \in \mathbb{C}^*. These equivalence classes, often called rays or one-dimensional subspaces, form the projective Hilbert space \mathbb{P}(H), which generalizes the finite-dimensional complex projective space \mathbb{CP}^n to infinite dimensions. This structure arises naturally in the foundations of , where pure quantum states are represented not by vectors in H but by rays in \mathbb{P}(H), as overall normalization and global phase factors have no physical significance and do not affect measurement probabilities via the . The space \mathbb{P}(H) inherits a natural geometry from H, including a Fubini-Study that induces a Kähler structure, enabling the formulation of as flows on this manifold. In infinite dimensions, \mathbb{P}(H) requires careful definition of topologies—such as the quotient topology or weak topologies—to ensure measurability and continuity of quantum operations, addressing challenges absent in finite-dimensional cases. Key applications include , where \mathbb{P}(H) serves as a for classical-like descriptions of quantum evolution, and the study of projective unitary representations of symmetry groups in .

Mathematical Foundations

Hilbert Spaces and Rays

A is defined as a complete over the complex numbers, meaning it is a equipped with an inner product that induces a , and every converges within the space. In the context of , Hilbert spaces are typically taken to be separable, ensuring a countable exists, which facilitates the representation of physical systems with discrete and continuous spectra. In , the state of a is represented by a ψ in the , subject to the condition ‖ψ‖ = 1, where the norm is derived from the inner product ⟨ψ|ψ⟩ = 1. This ensures the lies on the unit sphere in the , but physical observables depend only on directions rather than specific representatives. A in the is the one-dimensional subspace generated by a non-zero ψ, consisting of all scalar multiples of the form { c \psi \mid c \in \mathbb{C} \setminus {0} }. In , these rays capture the physically equivalent states, as overall scaling and global phase factors have no physical significance. For example, in the finite-dimensional ℂ^n, rays correspond to lines through the origin, with each ray intersecting the unit sphere S^{2n-1} in a circle, representing distinct directions modulo phase and scaling. were formalized by in the 1920s and 1930s as the mathematical foundation for , building on earlier work by and others to rigorously describe infinite-dimensional systems. The , to be detailed later, emerges as the of these rays under equivalence.

Equivalence Relations and Quotient Construction

The projective Hilbert space arises from the H through an that identifies vectors differing only by a non-zero scalar, reflecting the physical indistinguishability of quantum states under \mathbb{C}^* transformations. Specifically, two nonzero vectors \psi, \phi \in H \setminus \{0\} are equivalent, denoted \psi \sim \phi, if there exists \lambda \in \mathbb{C}^* such that \phi = \lambda \psi. This relation partitions H \setminus \{0\} into equivalence classes, each corresponding to a one-dimensional or "" in H. The projective , denoted \mathrm{PH}(H), is then defined as the set \mathrm{PH}(H) = (H \setminus \{0\}) / \sim, where elements are equivalence classes [\psi] = \{\lambda \psi \mid \lambda \in \mathbb{C}^*\}. An equivalent construction focuses on the unit sphere S(H) = \{\psi \in H \mid \|\psi\| = 1\} in H, which is a complete metric space under the norm topology. The group U(1) acts freely and continuously on S(H) via multiplication by unit complex scalars, and \mathrm{PH}(H) is the orbit space S(H) / U(1), inheriting the quotient topology. In this topology, a set U \subset \mathrm{PH}(H) is open if its preimage under the quotient map \pi: S(H) \to \mathrm{PH}(H), defined by \pi(\psi) = [\psi], is open in S(H). This ensures that \mathrm{PH}(H) is a topological space where convergence of sequences [\psi_n] \to [\psi] corresponds to the existence of phases \lambda_n \in U(1) such that \lambda_n \psi_n \to \psi in the norm of H. For a separable infinite-dimensional Hilbert space H, the quotient \mathrm{PH}(H) is Hausdorff and well-behaved topologically. The Hausdorff property follows from the compactness of U(1), which implies that the quotient map \pi separates distinct orbits: for distinct [\psi] and [\phi], there exist disjoint open neighborhoods in S(H) that are saturated under the U(1)-action, projecting to disjoint opens in \mathrm{PH}(H). Moreover, since S(H) is second-countable (as H is separable), the quotient topology on \mathrm{PH}(H) is also second-countable, making it a metrizable space without pathologies like non-separated points. This construction yields a locally compact Hausdorff space that serves as the state space in quantum mechanics. Unlike the real projective space \mathbb{RP}^n, which quotients \mathbb{R}^{n+1} \setminus \{0\} by positive \mathbb{R}^+, the structure of \mathrm{PH}(H) involves the \mathbb{C}^* rather than \mathbb{R}^*, leading to a with distinct geometric properties such as dimension and Kähler potential (though the is addressed elsewhere). This emphasis on non-zero scalars ensures that \mathrm{PH}(H) captures the invariance inherent to quantum superpositions, distinguishing it from real projective geometries.

Geometric and Topological Structure

Projective Space as a Manifold

In the finite-dimensional case, where the Hilbert space H = \mathbb{C}^n is an n-dimensional complex vector space, the projective Hilbert space \mathrm{PH}(H) is isomorphic to the \mathbb{C}\mathbb{P}^{n-1}, which has complex dimension n-1. This space consists of equivalence classes of nonzero vectors in H, where two vectors z, z' \in H \setminus \{0\} are identified if z' = \lambda z for some \lambda \in \mathbb{C}^* = \mathbb{C} \setminus \{0\}. Points in \mathbb{C}\mathbb{P}^{n-1} are represented using homogeneous coordinates [z_1 : z_2 : \dots : z_n], where (z_1, \dots, z_n) \in \mathbb{C}^n \setminus \{0\} and the notation indicates equivalence under \mathbb{C}^*-scaling. To endow \mathbb{C}\mathbb{P}^{n-1} with a manifold structure, it is covered by n standard affine charts U_k = \{ \in \mathbb{C}\mathbb{P}^{n-1} \mid z_k \neq 0 \} for k = 1, \dots, n, each diffeomorphic to \mathbb{C}^{n-1}. On U_k, the holomorphic coordinates are given by w_i^{(k)} = z_i / z_k for i \neq k, providing a local identification with \mathbb{C}^{n-1}. The overlaps between charts ensure compatibility through transition functions that are holomorphic. For instance, on U_j \cap U_k with j \neq k, the transition from coordinates on U_j to those on U_k is w_i^{(k)} = w_i^{(j)} / w_j^{(j)} for i \neq k, j, while the coordinate corresponding to the j-th direction in the U_k-system is w_j^{(k)} = 1 / w_j^{(j)}. These transition maps are rational functions that are holomorphic on the overlaps, confirming that the atlas \{ (U_k, \phi_k) \mid k = 1, \dots, n \}, where \phi_k denotes the coordinate map, defines a structure on \mathbb{C}\mathbb{P}^{n-1}. In the infinite-dimensional setting, for a separable infinite-dimensional complex H, the \mathrm{PH}(H) inherits a smooth manifold structure as the of the unit in H by the S^1-, resulting in a non-compact infinite-dimensional . Unlike the finite-dimensional case, it lacks a finite atlas of charts analogous to the affine coverings, but it supports a differentiable structure compatible with the .

Fubini-Study Metric and Kähler Geometry

The Fubini-Study on the projective \mathbb{P}(\mathcal{H}), where \mathcal{H} is a finite-dimensional of dimension n+1, arises as the quotient induced from the round on the unit sphere S^{2n+1} \subset \mathcal{H} via the \pi: S^{2n+1} \to \mathbb{C}\mathbb{P}^n. Specifically, for tangent vectors H_1, H_2 at a line L \in \mathbb{C}\mathbb{P}^n, the is given by g_{\mathrm{FS},L}(H_1, H_2) = g_{S_x}(H_1 x, H_2 x), where x \in L \cap S^{2n+1} and g_S is the standard round on the sphere, ensuring the projection is a Riemannian submersion. In affine charts, such as the standard chart on \mathbb{C}^n where [z_0 : z_1 : \cdots : z_n] = [1 : w_1 : \cdots : w_n] with w = (w_1, \dots, w_n), the Fubini-Study metric takes the local form ds^2 = \frac{\sum_{i=1}^n dw_i d\bar{w}_i}{(1 + \|w\|^2)^2}, which is the Hermitian metric g_{i\bar{j}} = \frac{\delta_{ij} (1 + \|w\|^2) - \bar{w}_i w_j}{(1 + \|w\|^2)^2}. This expression derives from the Kähler potential K(w) = \log(1 + \|w\|^2) in these coordinates, where the metric components are the mixed partial derivatives g_{i\bar{j}} = \partial_i \partial_{\bar{j}} K. In homogeneous coordinates [Z] = [Z_0 : \cdots : Z_n] with Z \in \mathcal{H} \setminus \{0\}, the Kähler potential is K(Z) = \log \sum_{i=0}^n |Z_i|^2, providing a gauge-invariant description that descends to the quotient. The associated Kähler form is the fundamental 2-form \omega = \frac{i}{2} \partial \bar{\partial} \log(1 + \|w\|^2) in affine coordinates, or more generally \omega = \frac{i}{2} \partial \bar{\partial} K, making \mathbb{C}\mathbb{P}^n a . This form is closed (d\omega = 0) and compatible with the complex structure, endowing the space with a natural structure where \omega serves as the symplectic 2-form. Consequently, the Fubini-Study geometry supports , with geodesics corresponding to Hamiltonian flows generated by the metric's symplectic potential. The Fubini-Study metric exhibits constant holomorphic sectional curvature equal to 4 (in the standard normalization), with sectional curvatures ranging from 1 to 4, achieving the maximum on holomorphic planes. It is an Einstein metric, satisfying \mathrm{Ric}(g_{\mathrm{FS}}) = 2(n+1) \, g_{\mathrm{FS}} for \dim_{\mathbb{C}} \mathbb{C}\mathbb{P}^n = n, and thus Kähler-Einstein, reflecting its role as a of rank 1. In the infinite-dimensional case, where \mathcal{H} is a separable complex , the Fubini-Study extends to \mathbb{P}(\mathcal{H}) via the same quotient construction from the unit sphere in \mathcal{H}, yielding a complete Kähler \sigma^2 that generalizes the finite-dimensional version. However, challenges arise in ensuring well-definedness and completeness, often requiring weak topologies on \mathcal{H} to handle of rays and to embed bounded domains isometrically into \mathbb{P}(\mathcal{H}) using Bergman kernels.

Properties and Theorems

Dimension and Homogeneity

The projective Hilbert space PH(H) associated to a finite-dimensional complex Hilbert space H of dimension n is diffeomorphic to the \mathbb{CP}^{n-1}, a smooth manifold of real dimension $2(n-1). In the infinite-dimensional case, where \dim H = \infty, PH(H) has infinite real dimension and is non-compact. For finite n, \mathbb{CP}^{n-1} is compact as a quotient of the compact unit sphere in H by the S^1-action. Its singular homology groups satisfy H_k(\mathbb{CP}^{n-1}; \mathbb{Z}) = \mathbb{Z} for k = 0, 2, \dots, 2(n-1) and H_k(\mathbb{CP}^{n-1}; \mathbb{Z}) = 0 otherwise, reflecting its CW-complex structure with one cell in each even dimension from 0 to $2(n-1). The Euler characteristic is \chi(\mathbb{CP}^{n-1}) = n, obtained as the alternating sum of Betti numbers or the number of CW-cells. As a homogeneous space, \mathbb{CP}^{n-1} \cong U(n) / (U(1) \times U(n-1)), where the U(n) acts transitively by transforming lines in H into each other. The integral cohomology ring is H^*(\mathbb{CP}^{n-1}; \mathbb{Z}) \cong \mathbb{Z} / (x^n), where x is the generator in degree 2 represented by the Kähler class [\omega] of the Fubini-Study form.

Unitary Group Action

The unitary group U(\mathcal{H}) acts on the Hilbert space \mathcal{H} by unitary operators, which preserve the inner product and thus map rays to rays, inducing an action on the projective Hilbert space \mathbb{P}\mathcal{H} defined by [U\psi] = [U \psi], where rays are preserved up to a global phase factor e^{i\theta} with \theta \in \mathbb{R}. This action is well-defined because unitary operators maintain the equivalence relation on nonzero vectors, ensuring that the projective space, consisting of one-dimensional subspaces, is invariant under the transformation. The kernel of this induced action is the center U(1), consisting of phase multiplications, so the effective group acting on \mathbb{P}\mathcal{H} is the projective unitary group \mathrm{PU}(\mathcal{H}) = U(\mathcal{H}) / U(1). For finite-dimensional \mathcal{H} = \mathbb{C}^n, this is \mathrm{PU}(n), which acts transitively and effectively on \mathbb{P}\mathbb{C}^n \cong \mathbb{CP}^{n-1}, meaning any two rays can be mapped to each other by some element of \mathrm{PU}(n), and the action is faithful. The stabilizer of a specific point, such as the ray [e_1] spanned by the first standard basis vector, is the subgroup isomorphic to U(n-1), consisting of unitaries that fix this ray up to phase. Under the action of \mathrm{PU}(n), the entire space \mathbb{CP}^{n-1} forms a single , confirming its homogeneity as a manifold where the realizes the transitive . In the infinite-dimensional case for separable \mathcal{H}, \mathrm{PU}(\mathcal{H}) similarly acts transitively on \mathbb{P}\mathcal{H}, with the projective space being the set of all one-dimensional subspaces. The infinitesimal action arises from the \mathfrak{u}(n), consisting of skew-Hermitian matrices, which generates derivations on \mathbb{CP}^{n-1} via the fundamental vector fields; these correspond to Killing vector fields with respect to the Fubini-Study , where the is proportional to the negative restricted to the . As a consequence, \mathbb{P}\mathcal{H} is a symmetric space under the action of \mathrm{PU}(\mathcal{H}), characterized by an involution at each point whose fixed-point set is the isotropy subgroup, yielding a Riemannian symmetric structure invariant under the group.

Applications in Physics

Quantum Mechanics State Space

In quantum mechanics, the physical states of a system are represented by rays in the Hilbert space rather than individual vectors, owing to the invariance of observables under global phase transformations. Specifically, for a normalized state vector |\psi\rangle \in \mathcal{H}, the expectation value of an operator A is \langle \psi | A | \psi \rangle, which remains unchanged if |\psi\rangle is replaced by e^{i\theta} |\psi\rangle for any real \theta. This phase invariance implies that states differing only by a phase factor describe the same physical situation, leading to the identification of equivalence classes known as rays [\psi], which form the projective Hilbert space \mathbb{P}(\mathcal{H}). Pure quantum states correspond bijectively to rank-1 orthogonal projectors P_\psi = |\psi\rangle\langle\psi|, where the density operator for the pure state is \rho = P_\psi, and the expectation value of an A is given by \operatorname{Tr}(\rho A). This representation ensures that the projective structure captures the physically distinguishable states without redundancy from phase factors. Superpositions in the \mathcal{H} map to nonlinear combinations in \mathbb{P}(\mathcal{H}), as the projection operation normalizes linear combinations, which underpins quantum phenomena essential to the theory. A canonical example is the of a single , which is the \mathbb{CP}^1 \cong S^2, known as the . Basis states such as |0\rangle correspond to the , while equatorial points represent equal superpositions like \frac{1}{\sqrt{2}}(|0\rangle + |1\rangle). The governs measurement probabilities: for normalized states |\phi\rangle and |\psi\rangle, the probability of outcome associated with |\phi\rangle when measuring in the basis including |\psi\rangle is | \langle \phi | \psi \rangle |^2, which in the projective space is the squared of the overlap between rays [ \phi ] and [ \psi ]; this overlap relates to the Fubini-Study measuring state distinguishability. Historically, Paul Dirac's introduction of bra-ket notation in the 1930s implicitly incorporated the projective nature of quantum states by treating kets as abstract vectors in while emphasizing phase-invariant inner products.

Representation in Phase Space Formulations

In phase space formulations of , coherent states for bosonic systems offer a natural parameterization of the projective PH(H). These states, labeled by complex parameters α ∈ ℂ^n, correspond to points in the classical , parameterized by points in \mathbb{C}^n \cong \mathbb{R}^{2n}, the classical for n modes, via coherent states labeled by \alpha \in \mathbb{C}^n. Each coherent state |α⟩ is a minimum-uncertainty Gaussian displaced in , minimizing the Heisenberg uncertainty and bridging quantum and classical descriptions. The set of coherent states forms an overcomplete basis in the , enabling a resolution of the : \int_{\mathbb{C}^n} |\alpha\rangle \langle \alpha| \frac{d^2\alpha}{\pi^n} = I, where dμ(α) = d²α / π^n is the invariant measure on ℂ^n, ensuring that any state can be expanded in this basis despite the redundancy. This overcompleteness facilitates phase space integrals over the , useful for computing expectation values and evolving states under bosonic Hamiltonians. The Husimi Q-function provides a smoothed quasi-probability on this , defined for a operator ρ as Q(\alpha) = \frac{1}{\pi^n} \langle \alpha | \rho | \alpha \rangle. It represents quantum states on PH(H) as non-negative functions, convolving the Wigner function with a Gaussian to avoid negative values, and integrates to unity over the . This is particularly valuable for visualizing quantum interference and classical-like behavior in bosonic systems. The projective Hilbert space relates to as the quantum analog of classical structures, such as the T^*ℝ^n modulo ℤ_2 identifications, where the Fubini-Study symplectic form underlies the quantization procedure. The Bargmann transform further illuminates this by unitarily mapping L²(ℝ^n) to the of holomorphic functions on ℂ^n, equipped with a that preserves the Kähler structure and reveals the of the . A canonical example is the quantum harmonic oscillator, where the infinite-dimensional Hilbert space yields PH(H) parameterized by the complex plane ℂ ≅ ℝ², the classical phase space of position and momentum. Here, coherent states |α⟩ trace circular orbits under time evolution, mirroring classical trajectories while incorporating quantum fluctuations via the Q-function.

Generalizations and Extensions

Infinite-Dimensional Cases

In the infinite-dimensional case, the projective Hilbert space PH(H) for a separable Hilbert space H, such as ℓ²(ℕ), differs significantly from its finite-dimensional analogue due to topological and geometric pathologies. While H admits a countable orthonormal basis, PH(H) is non-compact and lacks the smooth manifold structure of finite-dimensional complex projective spaces, as it cannot be covered by a single chart or finite atlas in a Banach manifold sense; instead, it is constructed as the inductive limit of the finite-dimensional projective spaces PH(H_n), where H_n is the n-dimensional subspace spanned by the first n basis vectors. This inductive limit topology ensures that PH(H) is paracompact and metrizable via a Fubini-Study-like metric defined locally on each PH(H_n), but the global structure avoids the pitfalls of attempting a uniform infinite-dimensional metric. The Fubini-Study on PH() is defined using the Kähler potential ∂∂̅ log‖Z‖² on affine charts, extending the finite-dimensional formula, but its global implementation requires care due to the non-compactness of the in . PH() is separable as a under this (or equivalent ones like the ρ([φ],[ψ]) = √(1 - |⟨φ|ψ⟩|² for vectors φ, ψ), coinciding with the induced by transition probabilities tr(PQ) for rank-one projectors P, Q, making it a second-countable . However, convergence in PH() often involves subtleties with operator : sequences of projectors converge in the strong operator if ‖P_n x - P x‖ → 0 for all x ∈ , whereas the weak operator uses ⟨P_n x, y⟩ → ⟨P x, y⟩ for all x, y ∈ ; for pure states, these align with the on PH(), but distinctions arise in unbounded settings. Generalized coherent states can be constructed such that their is dense in H, implying that their corresponding rays densely span PH(H) in the weak or metric topology. These states, constructed via group actions or from finite-dimensional cases, provide a practical way to approximate any pure state, leveraging overcompleteness without a resolution of the in the infinite-dimensional limit. Elements of PH(H) are identified with rank-one orthogonal projectors (projective units), facilitating this density. In applications to (QFT), PH(H) serves as the state space for the of bosonic or fermionic fields, which is separable and infinite-dimensional, but practical computations face challenges due to ultraviolet divergences that prevent direct use of the full PH(H) without regularizing infinite volumes or modes. Unlike the finite-dimensional case, the projective PU(H) = U(H)/U(1) lacks a bi-invariant , as U(H) is not locally compact; instead, invariant means or approximate measures on finite-dimensional approximations are employed for integration over state spaces or representations. This absence underscores the need for inductive or weak topologies in infinite-dimensional quantum theories to maintain mathematical rigor.

Analogues in Other Settings

The real projective Hilbert space arises as the projectivization of a real Hilbert space H, defined as \mathbb{RP}(H) = (H \setminus \{0\}) / (\mathbb{R} \setminus \{0\}), where \mathbb{R} \setminus \{0\} acts by non-zero scalar multiplication. For a finite-dimensional real Hilbert space of dimension n, this space is the real projective space \mathbb{RP}^{n-1} of real dimension n-1}. In classical mechanics, real projective Hilbert spaces appear in formulations of projective dynamics, where properties like the closure of Keplerian orbits after one period are invariant under projective transformations. Grassmannians provide higher-rank analogues of projective Hilbert spaces, generalizing the space of 1-dimensional subspaces to the space of k-dimensional subspaces in a H. The \mathrm{Gr}_k(H) consists of equivalence classes of k-planes in H, embedding into a via the , which realizes it as a projective algebraic variety. For k=1, \mathrm{Gr}_1(H) recovers the projective Hilbert space, highlighting its role as the foundational case in this hierarchy. In , introduced by in the 1960s, the projective twistor space \mathbb{PT} serves as an analogue of projective Hilbert space, modeling aspects of geometry. Specifically, \mathbb{PT} \cong \mathbb{CP}^3 is the projectivization of the \mathbb{T} = \mathbb{C}^4, establishing a correspondence between null geodesics in complexified Minkowski and points in \mathbb{PT}. Non-commutative geometry extends the projective Hilbert space framework through spectral triples (A, H, D), where A is a non-commutative algebra on a H, and D is a Dirac-like encoding information. In this setting, classical vector bundles are replaced by finitely generated projective modules over A, providing a non-commutative analogue of projective spaces that captures geometric structures on "non-commutative manifolds." Projective spinor bundles in generalize projective structures to spinor contexts, particularly in relation to on Riemannian manifolds. For a spin manifold M, the projective spinor bundle is constructed as \mathbb{P}^\pm(M) = S(M) \times_{\sigma^\pm} \mathbb{P}(\Delta^\pm_4), where S(M) is the and \Delta^\pm_4 are the half-spin representations; this structure admits a compatible whose symbol induces the 's irreducibility. The infinite-dimensional projective Hilbert space \mathbb{PH}(H) possesses a universal property as the classifying space for U(1)-principal bundles (or equivalently, complex line bundles), where homotopy classes of maps from a space X to \mathbb{PH}(H) \cong \mathbb{CP}^\infty classify such bundles over X.

References

  1. [1]
    [PDF] 2 Hilbert Space
    2.1 Definition of Hilbert Space. Hilbert space is a vector space H over C that is equipped with a complete inner product.
  2. [2]
    [PDF] B Basic concepts from quantum theory - UTK-EECS
    Projective Hilbert space: Pure states correspond to the rays in a projective Hilbert space. A ray is an equivalence class of nonzero vectors under the ...
  3. [3]
    The Projective Hilbert Space as a Classical Phase ... - ResearchGate
    Aug 7, 2025 · The projective Hilbert space carries a natural symplectic structure which enables one to reformulate quantum dynamics as a classical Hamiltonian ...
  4. [4]
  5. [5]
    [PDF] 2006 Lecture Notes on Hilbert Spaces and Quantum Mechanics
    Inspired by quantum mechan- ics (see below), in that year von Neumann gave the definition of a Hilbert space as an abstract mathematical structure, as ...
  6. [6]
    [PDF] Hilbert Space Quantum Mechanics
    ◦ The term projective decomposition of the identity is sometimes used to distinguish what we are talking about from a POVM (positive operator-valued measure), ...
  7. [7]
    [PDF] Hilbert Space Quantum Mechanics
    There is therefore a one-to-one correspondence between directions, or the corresponding points on the unit sphere, with rays of a two-dimensional Hilbert space.
  8. [8]
    Geometry of Quantum States
    Geometry of Quantum States. An Introduction to Quantum Entanglement. Search within full ...
  9. [9]
    Topologies and Measurable Structures on the Projective Hilbert Space
    A systematic review of the various topologies that can be defined on the projective Hilbert space P(H), ie, on the set of the pure quantum states, is presented.
  10. [10]
    [PDF] Projective spaces, the Fubini-Study metric and a little bit more
    The projective space (PV) of V is defined as {L ⊆ V | L is a subspace of V with dim L = 1}, or (V \{0})/∼.
  11. [11]
    [PDF] Topologies and Measurable Structures on the Projective Hilbert Space
    Aug 9, 2007 · Abstract. A systematic review of the various topologies that can be defined on the projective Hilbert space ... Infinite-Dimensional Hamiltonian ...
  12. [12]
    [PDF] Lectures on Kähler Geometry - HAL
    Feb 13, 2004 · The Fubini–Study metric was defined via its fundamental 2–form, which was expressed by local Kähler potentials. We provide here a more ...
  13. [13]
    GEOMETRY OF BOUNDED DOMAINS
    SHOSHICHI kobayashi. In this paper, we shall study differential geometric properties of bounded domains in Cn. Here is the summary of our results.
  14. [14]
    [PDF] Complex Projective Space Definition
    Complex projective n-space, denoted by CPn, is defined to be the set of 1-dimensional complex-linear subspaces of Cn+1, with the quotient topology inherited ...
  15. [15]
    An infinite dimensional balanced embedding problem I:existence
    Sep 3, 2023 · We investigate the problem of balanced embedding of a non-compact complex manifold into an infinite-dimensional projective space.
  16. [16]
    [PDF] Algebraic Topology I: Lecture 16 Homology of CW-Complexes
    Complex projective space CPn has a CW structure in which. Sk2kCPn = Sk2k+1CPn = CPk . The attaching S2k−1 → CPk sends v ∈ S2k−1 ⊆ Cn to the complex line through ...
  17. [17]
    [PDF] Class 19. Examples of Kähler manifolds (November 5) The Hodge ...
    Nov 5, 2024 · Its cohomology is easy to compute, using some results from algebraic topology. Lemma 19.2. The cohomology groups of complex projective space are.
  18. [18]
    [PDF] GEOMETRY OF QUANTUM STATES
    The geometry of quantum states is a highly interesting subject in itself ... Hilbert space. In this case we can take the word geometry literally: there ...
  19. [19]
    [PDF] arXiv:1508.07209v3 [math-ph] 3 May 2016
    May 3, 2016 · In the case of P(H) the isomtery group is the projective unitary group PU(n) then we have dim(Kill(P(H))) = n2 − 1.
  20. [20]
  21. [21]
    Coherent and Incoherent States of the Radiation Field | Phys. Rev.
    Coherent states reduce field correlation functions to factorized forms, while incoherent fields are generated by superposing outputs of many stationary sources.Missing: original | Show results with:original
  22. [22]
    [PDF] Coherent States - arXiv
    Mar 29, 2009 · Coherent states (of the harmonic oscillator) were introduced by Erwin Schrödinger. (1887-1961) at the very beginning of quantum mechanics in ...Missing: original | Show results with:original
  23. [23]
    Husimi's Q(α) function and quantum interference in phase space
    Mar 10, 2004 · Husimi's Q(α) function and quantum interference in phase space. D F Mundarain and J Stephany. Published 10 March 2004 • 2004 IOP Publishing ...Missing: original paper
  24. [24]
    The Projective Hilbert Space as a Classical Phase Space for ...
    The projective Hilbert space carries a natural symplectic structure which enables one to reformulate quantum dynamics as a classical Hamiltonian one.
  25. [25]
    Fubini-Study metric for an infinite dimensional Hilbert space
    Mar 30, 2013 · Can one define a Fubini-Study metric/Kaehler metric on the projective space of an infinite dimensional Hilbert space, ie using the formula ∂ˉ∂log|Z|2?Isometry group of the Fubini-Study metric on complex projective ...Explicit construction of Fubini Study Metric - MathOverflowMore results from mathoverflow.net
  26. [26]
    infinitesimal neighborhoods of infinite-dimensional complex ...
    Let V be an infinite dimensional complex. Banach space and Y a complex Banach manifold containing. X := P(V ) as a codimension r closed split submanifold.
  27. [27]
    [PDF] Coherent States and Fubini-Study Geometry Timothy R. Field - People
    The metric on the ambient state space is the familiar Fubini-Study metric, and we use this to calculate the induced metric on C. There are some technicalities ...
  28. [28]
    projective unitary group in nLab
    Sep 19, 2021 · For ℋ \mathcal{H} a Hilbert space, the projective unitary group P U ( ℋ ) P U(\mathcal{H}) is the quotient of the unitary group U ( ℋ ) U(\ ...
  29. [29]
    [PDF] Information geometry in vapour-liquid equilibrium - arXiv
    Sep 6, 2008 · the real projective Hilbert space. Suppose now that we consider the Einstein equation. (44) for the metric on the projective Hilbert space ...
  30. [30]
  31. [31]
    Projective dynamics and classical gravitation | Regular and Chaotic ...
    Dec 16, 2008 · We show that two properties in classical particle dynamics are projective properties. The fact that the Keplerian orbits close after one turn is ...
  32. [32]
    [PDF] Cohomology of the complex Grassmannian
    The Grassmannian is a generalization of projective spaces–instead of looking at the set of lines of some vector space, we look at the set of all n-planes ...Missing: higher- rank
  33. [33]
    [PDF] Projective Spaces, Grassmannians and the Plücker Embedding
    ... projective spaces which are sets of one- dimensional subspaces. This concept is then generalized to higher dimensions, in the chapter on Grassmannians.
  34. [34]
    Twistor Algebra | Journal of Mathematical Physics - AIP Publishing
    Penrose; Twistor Algebra. J. Math. Phys. 1 February 1967; 8 (2): 345–366 ... This content is only available via PDF. Open the PDF for in another window.
  35. [35]
  36. [36]