Fact-checked by Grok 2 weeks ago

Fock space

In , Fock space is a that describes the quantum states of systems with a variable or indeterminate number of identical particles, enabling the representation of states across different particle numbers within a single framework. It is constructed as the direct sum \mathcal{F}(H) = \bigoplus_{n=0}^\infty H^{\otimes n}, where H is the single-particle , with the n=0 term being the vacuum state and higher terms symmetrized for bosons or antisymmetrized for fermions to account for indistinguishability and statistics. This structure facilitates the use of , which obey canonical commutation relations [a_i, a_j^\dagger] = \delta_{ij} for bosons or anticommutation relations \{b_i, b_j^\dagger\} = \delta_{ij} for fermions, allowing precise counting of occupation numbers in basis states. Named after Soviet physicist , who introduced the concept in 1932, Fock space emerged as a key tool in to handle many-body problems beyond fixed particle numbers, such as in and . For bosonic systems like photons, the symmetric Fock space \Gamma_s(H) permits unlimited occupation of states (n_i \geq 0), embodying Bose-Einstein statistics, while the fermionic antisymmetric Fock space \Gamma_a(H) enforces the with at most one particle per state (n_i \in \{0,1\}). These spaces underpin the formalism for quantum gases, superconductors, and relativistic fields, where the serves as the annihilated by all destruction operators.

Basic Concepts

Single-particle Hilbert space

The single-particle \mathcal{H}_1 is defined as a separable , meaning it possesses a and is complete with respect to the norm induced by its inner product, providing the mathematical framework for describing the quantum states of an individual particle over its configuration space. This separability ensures that \mathcal{H}_1 can accommodate infinite-dimensional representations typical in , allowing for a dense of basis vectors that span the space. In the context of building more complex , \mathcal{H}_1 forms the essential prerequisite structure. For non-relativistic particles moving in three-dimensional , \mathcal{H}_1 is commonly realized as L^2(\mathbb{R}^3), the space of square-integrable complex-valued functions \psi(\mathbf{r}) with respect to the , where the norm \|\psi\|^2 = \int_{\mathbb{R}^3} |\psi(\mathbf{r})|^2 d^3\mathbf{r} < \infty ensures physical wave functions are normalizable. In relativistic settings, such as for particles obeying the Klein-Gordon or , \mathcal{H}_1 is often formulated in momentum space to incorporate Lorentz invariance, taking the form of L^2(\mathbb{R}^3, d^3\mathbf{p}/(2\omega_{\mathbf{p}})) with \omega_{\mathbf{p}} = \sqrt{\mathbf{p}^2 + m^2}, where the measure accounts for the relativistic energy-momentum relation on the mass shell. The inner product on \mathcal{H}_1 is \langle \psi | \phi \rangle = \int \psi^*(\mathbf{r}) \phi(\mathbf{r}) d^3\mathbf{r} in the representation, enabling the computation of probabilities and overlaps between states. Basis states, such as eigenstates |\mathbf{r}\rangle or eigenstates |\mathbf{p}\rangle, are orthonormal in the distributional sense: \langle \mathbf{r} | \mathbf{r}' \rangle = \delta^3(\mathbf{r} - \mathbf{r}') and \langle \mathbf{p} | \mathbf{p}' \rangle = \delta^3(\mathbf{p} - \mathbf{p}'), reflecting the continuous of these observables despite the states not being square-integrable themselves. These generalized eigenstates form a complete set, allowing any state in \mathcal{H}_1 to be expanded as \psi(\mathbf{r}) = \langle \mathbf{r} | \psi \rangle or in momentum space via . As the foundational space, \mathcal{H}_1 underpins the construction of many-particle Hilbert spaces through tensor products, where the state of N distinguishable particles occupies \mathcal{H}_1^{\otimes N}, facilitating the description of interactions and correlations in multi-particle quantum systems. Fock space extends this to accommodate variable particle numbers via a over such tensor products.

Many-particle states for distinguishable particles

For a system of N distinguishable particles, each described by the single-particle Hilbert space \mathcal{H}_1, the many-particle Hilbert space \mathcal{H}_N is constructed as the tensor product \mathcal{H}_N = \mathcal{H}_1^{\otimes N}. This structure allows the state of the system to be specified independently for each particle, reflecting their distinguishability. The dimension of \mathcal{H}_N is the Nth power of the dimension of \mathcal{H}_1, enabling a basis expansion with N-fold products of single-particle basis states. States in \mathcal{H}_N are represented by wave functions \psi(x_1, x_2, \dots, x_N), where x_i denotes the coordinates (position, spin, etc.) of the ith particle. The inner product between two such states \psi and \phi is given by the integral over all variables: \langle \psi | \phi \rangle = \int \psi^*(x_1, \dots, x_N) \phi(x_1, \dots, x_N) \, dx_1 \dots dx_N, assuming the single-particle space is L^2 with respect to the appropriate measure. This ensures the Hilbert space is complete and equipped with a positive-definite inner product. A simple example is the two-particle state formed as a product of orthonormal single-particle states \phi_a(x_1) and \phi_b(x_2), yielding \psi(x_1, x_2) = \phi_a(x_1) \phi_b(x_2). For normalization, the condition \int |\psi(x_1, x_2)|^2 \, dx_1 dx_2 = 1 holds directly if \int |\phi_a(x_1)|^2 \, dx_1 = 1 and \int |\phi_b(x_2)|^2 \, dx_2 = 1, due to the separability of the . Orthogonality follows similarly: for distinct pairs, say \phi_a, \phi_b and \phi_c, \phi_d, the inner product \int \psi^*(x_1, x_2) \chi(x_1, x_2) \, dx_1 dx_2 = 0 if a \neq c or b \neq d, leveraging the \int \phi_a^*(x) \phi_c(x) \, dx = \delta_{ac}.

Indistinguishable particles and symmetrization

In , systems of require that the many-particle wave functions be adapted to account for their identical nature, projecting onto specific symmetry subspaces of the full \mathcal{H}_N = \mathcal{H}^{\otimes N}, where \mathcal{H} is the single-particle . This adaptation arises from the symmetrization postulate, which mandates that the total of identical particles must transform either symmetrically (for bosons) or antisymmetrically (for fermions) under arbitrary permutations of particle labels. For identical bosons, which obey Bose-Einstein statistics, the wave function \psi(x_1, \dots, x_N) must be totally symmetric under particle exchange. The symmetrization S_B projects onto this symmetric subspace via S_B \psi(x_1, \dots, x_N) = \frac{1}{N!} \sum_{P \in S_N} P \psi(x_1, \dots, x_N), where S_N denotes the of all N! permutations P, and the P exchanges the coordinates accordingly. The resulting symmetric subspace \mathcal{H}_N^B consists of all states invariant under such permutations, ensuring no distinction between particles. In contrast, identical fermions, such as electrons, follow Fermi-Dirac statistics and require totally antisymmetric wave functions, as dictated by the , which prohibits two fermions from occupying the same . The antisymmetrization operator S_F is given by S_F \psi(x_1, \dots, x_N) = \frac{1}{N!} \sum_{P \in S_N} (-1)^{|P|} P \psi(x_1, \dots, x_N), where |P| is the of the (even or odd). This projects onto the antisymmetric \mathcal{H}_N^F, where exchanging any two particles introduces a minus sign. A fundamental consequence is the , originally formulated to explain atomic spectra. An explicit basis for \mathcal{H}_N^F is provided by Slater determinants, which construct antisymmetric states from single-particle orbitals \phi_i(x): \Psi(x_1, \dots, x_N) = \frac{1}{\sqrt{N!}} \det \begin{pmatrix} \phi_1(x_1) & \phi_1(x_2) & \cdots & \phi_1(x_N) \\ \phi_2(x_1) & \phi_2(x_2) & \cdots & \phi_2(x_N) \\ \vdots & \vdots & \ddots & \vdots \\ \phi_N(x_1) & \phi_N(x_2) & \cdots & \phi_N(x_N) \end{pmatrix}. This form automatically enforces antisymmetry and the exclusion principle, as the determinant vanishes if any two orbitals are identical.

Construction of Fock Space

Direct sum over particle numbers

The Fock space provides a unified Hilbert space framework for systems with an indefinite number of indistinguishable particles, accommodating superpositions across different particle numbers. It was originally introduced by Vladimir Fock in the context of second quantization to describe the configuration space of quantum systems. Formally, given a single-particle Hilbert space \mathcal{H}, the bosonic Fock space \mathcal{F}_s is defined as the direct sum \mathcal{F}_s = \bigoplus_{N=0}^\infty \mathcal{H}_N^s, where \mathcal{H}_0^s = \mathbb{C} is the one-dimensional vacuum sector, and for N \geq 1, \mathcal{H}_N^s denotes the Hilbert space of symmetrized N-fold tensor products of \mathcal{H}, i.e., the subspace of \mathcal{H}^{\otimes N} invariant under permutations of the factors. Similarly, the fermionic Fock space \mathcal{F}_a is \mathcal{F}_a = \bigoplus_{N=0}^\infty \mathcal{H}_N^a, with \mathcal{H}_N^a being the antisymmetrized N-fold tensor product subspace of \mathcal{H}^{\otimes N}. These constructions ensure that states respect the statistics of bosons (symmetric under exchange) or fermions (antisymmetric under exchange). A general element (state) of the Fock space takes the form \Psi = \bigoplus_{N=0}^\infty \Psi_N, where \Psi_N \in \mathcal{H}_N (with \mathcal{H}_N denoting either \mathcal{H}_N^s or \mathcal{H}_N^a) and \Psi_N = 0 for all but finitely many N in the algebraic direct sum; the full Fock space is the completion thereof with respect to the inner product. The inner product between two states \Psi = \bigoplus_{N=0}^\infty \Psi_N and \Phi = \bigoplus_{N=0}^\infty \Phi_N is defined sector-wise as \langle \Psi | \Phi \rangle = \sum_{N=0}^\infty \langle \Psi_N | \Phi_N \rangle_{\mathcal{H}_N}, where \langle \cdot | \cdot \rangle_{\mathcal{H}_N} is the inner product on \mathcal{H}_N. This orthogonal direct sum structure preserves the Hilbert space properties across sectors. When \mathcal{H} is a separable , the resulting Fock space \mathcal{F} (bosonic or fermionic) is also separable and complete, forming a whose basis can be constructed from an of \mathcal{H}. The separability arises from the countable of separable spaces, while completeness follows from the uniform boundedness of the sector norms in the completion process. As an illustrative example for bosons, consider a system where particles occupy discrete modes labeled by a basis of \mathcal{H}; states with a fixed particle number N, such as those specifying occupation numbers in each mode (summing to N), lie entirely within the sector \mathcal{H}_N^s.

Creation and annihilation operators

In the context of Fock space, which is constructed as a direct sum over Hilbert spaces of varying particle numbers, provide the dynamical framework for transitioning between these sectors by adding or removing particles. For bosonic particles, the creation operator a^\dagger(f) for a single-particle state f \in \mathcal{H}_1 acts on a Fock space state \Psi = \oplus_N \Psi_N by appending f to the N-particle component via the product, yielding a^\dagger(f) \Psi = \oplus_N \sqrt{N+1} (f \otimes \Psi_N), where the square root factor ensures normalization consistent with the bosonic commutation relations. The annihilation a(g) is the adjoint of the creation operator, satisfying a(g) = [a^\dagger(g)]^\dagger, and removes a particle in the direction of g \in \mathcal{H}_1. For fermionic particles, the creation operator a^\dagger(f) similarly appends f but uses the antisymmetric exterior product to enforce the , defined as a^\dagger(f) \Psi = \oplus_N (f \wedge \Psi_N), where \wedge denotes the product in the algebra. The corresponding annihilation operator a(g) is again the , with the fermionic nature imposing anticommutation relations that prevent double occupancy. The algebraic structure of these operators is governed by their commutation or anticommutation relations: for bosons, [a(f), a^\dagger(g)] = \langle f | g \rangle \mathbf{1}, where \mathbf{1} is the identity operator and the inner product \langle f | g \rangle projects onto the vacuum sector; for fermions, the anticommutator \{a(f), a^\dagger(g)\} = \langle f | g \rangle \mathbf{1} holds, with all other anticommutators vanishing. In a basis expansion, the field operators are expressed as linear combinations over an orthonormal basis \{\phi_k\} of the single-particle space, such that the creation field operator is \hat{\psi}^\dagger(x) = \sum_k \phi_k(x) a^\dagger_k, where a^\dagger_k = a^\dagger(\phi_k) creates a particle in the mode \phi_k at position x, and the annihilation field \hat{\psi}(x) is its adjoint.

Vacuum state and number operator

The vacuum state in Fock space, denoted |0\rangle, represents the absence of any particles and is the unique normalized vector in the zero-particle sector of the direct sum construction, expressed as |0\rangle = (1, 0, 0, \dots), where the leading 1 corresponds to the scalar state in the one-dimensional \mathcal{H}^{(0)} for zero particles, and all subsequent components in the n-particle sectors \mathcal{H}^{(n)} (n \geq 1) are zero vectors. This state satisfies the normalization condition \langle 0 | 0 \rangle = 1. It is annihilated by every annihilation operator acting on the space, such that a(f) |0\rangle = 0 for any single-particle wave function f in the underlying one-particle . The total number operator \hat{N}, which measures the overall particle count in a given state, is constructed from the . In a discrete { \phi_k } of the single-particle space, it takes the form \hat{N} = \sum_k a_k^\dagger a_k, where a_k and a_k^\dagger are the mode-specific annihilation and creation operators. Equivalently, in the continuous representation using field operators, \hat{N} = \int dx , \hat{\psi}^\dagger(x) \hat{\psi}(x), where \hat{\psi}(x) and \hat{\psi}^\dagger(x) annihilate and create particles at position x, respectively. Multi-particle states in Fock space are generated by successive applications of operators to the , and these states, labeled by occupation numbers \vec{n} = (n_1, n_2, \dots) with n_k denoting the number of particles in mode k, serve as eigenstates of \hat{N}: \hat{N} | \vec{n} \rangle = \left( \sum_k n_k \right) | \vec{n} \rangle. The eigenvalue \sum_k n_k gives the total particle number, and the normalization of such states follows from that of the , ensuring \langle \vec{n} | \vec{n} \rangle = 1 for properly symmetrized or antisymmetrized combinations appropriate to bosons or fermions.

Properties and Basis

Occupation number basis

The occupation number basis constitutes a canonical orthonormal basis for Fock space, consisting of states labeled by a vector \vec{n} = (n_1, n_2, \dots ), where each n_k denotes the number of particles occupying the k-th single-particle mode from a complete of the single-particle \mathcal{H}_1. This representation, introduced in the context of , facilitates the description of arbitrary many-particle states without fixing the particle number, allowing for superpositions across different particle sectors. For bosonic particles, the basis states are constructed as |\vec{n}\rangle = \prod_k \frac{(a^\dagger_k)^{n_k}}{\sqrt{n_k!}} |0\rangle, where a^\dagger_k is the creation operator for mode k, n_k = 0, 1, 2, \dots are non-negative integers, and |0\rangle is the state annihilated by all annihilation operators a_k. For fermionic particles, the restricts the occupation numbers to n_k = 0 or $1, yielding basis states of the form |\vec{n}\rangle = \prod_{k: n_k=1} a^\dagger_k |0\rangle, with the product taken over modes where particles are present; the fermionic anticommutation relations ensure automatic antisymmetrization. These states form an orthonormal set, satisfying \langle \vec{n} | \vec{m} \rangle = \delta_{\vec{n}\vec{m}}, and collectively span the entire Fock space \mathcal{F}. The completeness relation for the occupation number basis is \sum_{\vec{n}} |\vec{n}\rangle \langle \vec{n} | = I, where the sum runs over all possible occupation vectors \vec{n}, guaranteeing that any in \mathcal{F} can be expanded in this basis. This basis aligns with the direct-sum structure of Fock space over total particle numbers N = \sum_k n_k. The infinite-dimensional nature of \mathcal{F} arises from the unbounded possibilities for \vec{n} (unlimited for bosons, exponentially many configurations for fermions), but when \mathcal{H}_1 is separable, the basis ensures well-defined for operators and states.

Product states in Fock space

In Fock space, the basis states denoted by occupation number vectors \vec{n} = (n_1, n_2, \dots), where n_k represents the number of particles in the k-th single-particle orbital \phi_k, correspond to symmetrized product states of these orbitals when expressed in the position representation. These states provide a realization of the abstract number basis, ensuring proper accounting for particle indistinguishability by incorporating the required exchange symmetry. For identical bosons, the N-particle wave function with fixed occupations \sum_k n_k = N is the symmetrized product, mathematically expressed as a permanent: \psi_{\vec{n}}(\mathbf{x}_1, \dots, \mathbf{x}_N) = \frac{1}{\sqrt{\prod_k n_k!}} \sum_P \prod_k \prod_{j=1}^{n_k} \phi_k(\mathbf{x}_{P(j_k)}), where the sum runs over all distinct permutations P that assign the N particle coordinates to the occupied orbitals according to the multiplicities n_k. This form arises naturally in the configuration space formulation of for , guaranteeing full symmetry under particle exchange. For identical fermions, the corresponding basis states for a set of N singly occupied distinct orbitals \{\phi_1, \dots, \phi_N\} take the form of a : \psi_{\vec{n}}(\mathbf{x}_1, \dots, \mathbf{x}_N) = \frac{1}{\sqrt{N!}} \det \left[ \phi_j(\mathbf{x}_i) \right]_{i,j=1}^N. This antisymmetric construction enforces the , with each orbital occupied by at most one particle (n_k = 0 or $1), and was introduced as a practical method to satisfy fermionic symmetry in multi-electron systems. These symmetrized product representations serve as the fundamental "useful basis" for Fock space, as they eliminate overcounting of physically equivalent states that would occur in treatments of distinguishable particles by avoiding explicit particle labels. Instead, the occupation vector \vec{n} fully specifies the state, with permutations of coordinates leaving \psi_{\vec{n}} invariant due to the built-in symmetry. This uniqueness for fixed \vec{n} highlights the role of Fock space in modeling indistinguishable particles without redundancy.

Wave function representation

In Fock space, a general state is expressed as a linear superposition across sectors of different particle numbers N, where each N-particle component incorporates the appropriate symmetrization or antisymmetrization to account for the indistinguishability of particles. This representation bridges the operator formalism of with the intuitive picture from . The construction ensures that states with definite particle numbers are orthogonal, allowing for a natural description of systems with variable particle occupancy. The explicit form of such a state for particles is |\Psi\rangle = \sum_{N=0}^\infty \frac{1}{\sqrt{N!}} \int d\xi_1 \cdots d\xi_N \, \psi_N(\xi_1, \dots, \xi_N) \hat{\psi}^\dagger(\xi_1) \cdots \hat{\psi}^\dagger(\xi_N) |0\rangle, where \psi_N(\xi_1, \dots, \xi_N) is the N-particle , which must be fully symmetric for bosons or fully antisymmetric for fermions, \xi_i denote the single-particle coordinates (position, momentum, or other ), \hat{\psi}^\dagger(\xi) are the operators at those coordinates, and |0\rangle is the vacuum state. This integral form arises from expanding the state in a continuous basis of the single-particle , with the normalization factor \frac{1}{\sqrt{N!}} applying to both bosonic and fermionic cases, ensuring proper normalization consistent with commutation/anticommutation relations. The functions \psi_N(\xi_1, \dots, \xi_N) provide the for finding the system in a specific of N particles at positions \xi_1, \dots, \xi_N, interpreted probabilistically upon onto a fixed-N . The inner product between two general states |\Psi\rangle and |\Phi\rangle with corresponding wave functions \psi_N and \phi_N is \langle \Psi | \Phi \rangle = \sum_{N=0}^\infty \int \psi_N^*(\xi_1, \dots, \xi_N) \phi_N(\xi_1, \dots, \xi_N) \, d\xi_1 \cdots d\xi_N, which decomposes into independent contributions from each particle-number sector, guaranteeing across different N. In some formulations, particularly those emphasizing path integrals, the state can be viewed briefly as a functional \Psi[\hat{\psi}] over the field operators \hat{\psi}, facilitating connections to functional integral methods for computing . Product states, which fix a single N in the superposition, serve as of this general representation.

Applications

Second quantization in many-body physics

In many-body physics, reformulates the quantum mechanical description of interacting particles using Fock space, where states are labeled by occupation numbers rather than explicit coordinates of . This approach employs field operators \hat{\psi}^\dagger(x) and \hat{\psi}(x), which create and annihilate particles at position x, acting on the Fock space to naturally accommodate varying particle numbers N. The many-body is expressed as \hat{H} = \int dx \, \hat{\psi}^\dagger(x) h(x) \hat{\psi}(x) + \frac{1}{2} \iint dx \, dy \, \hat{\psi}^\dagger(x) \hat{\psi}^\dagger(y) V(x,y) \hat{\psi}(y) \hat{\psi}(x), where h(x) is the single-particle (typically including and external potentials) and V(x,y) represents the two-body interaction potential. This form arises from quantizing the classical field operators, ensuring antisymmetry for fermions or symmetry for bosons through the appropriate commutation relations. The single-particle term \int dx \, \hat{\psi}^\dagger(x) h(x) \hat{\psi}(x) captures non-interacting dynamics, such as the -\frac{\hbar^2}{2m} \nabla^2 for electrons, while the two-body term accounts for interactions like repulsion V(x,y) = \frac{e^2}{|x-y|}. In discrete models, such as those for , the adapts to site-based operators c^\dagger_{i\sigma} and c_{i\sigma}, where i denotes sites and \sigma , yielding a similar structure but with sums over sites instead of integrals. A canonical example is the , which describes strongly correlated electrons in narrow bands: \hat{H} = -t \sum_{\langle i,j \rangle, \sigma} \left( \hat{c}^\dagger_{i\sigma} \hat{c}_{j\sigma} + \hat{c}^\dagger_{j\sigma} \hat{c}_{i\sigma} \right) + U \sum_i \hat{n}_{i\uparrow} \hat{n}_{i\downarrow}, with hopping amplitude t, on-site repulsion U, and number operators \hat{n}_{i\sigma} = \hat{c}^\dagger_{i\sigma} \hat{c}_{i\sigma}. This Fock space representation facilitates studies of phenomena like Mott insulation and . The primary advantages of lie in its ability to handle particle interactions and indefinite N seamlessly, avoiding explicit symmetrization of wave functions. For instance, mean-field approximations like Hartree-Fock become straightforward, replacing the full interaction with effective single-particle potentials derived from expectation values in Fock space, enabling tractable computations for systems like electron gases. This framework also simplifies and diagrammatic expansions, making it indispensable for non-relativistic many-body problems in condensed matter.

Fock space in quantum field theory

In (QFT), the Fock space provides the for describing the quantum states of relativistic fields, incorporating both particle and while respecting Lorentz invariance and locality. For a free real of mass m, the single-particle \mathcal{H}_1 is constructed as the space of square-integrable wave functions over three-momentum space equipped with the Lorentz-invariant measure, given by \mathcal{H}_1 = L^2(\mathbb{R}^3, d^3k / ((2\pi)^3 2 \omega_k)), where \omega_k = \sqrt{|\vec{k}|^2 + m^2} is the relativistic energy. This space arises from the positive-frequency solutions to the Klein-Gordon equation on a Cauchy , ensuring that single-particle states transform irreducibly under the . The full Fock space \mathcal{F} is then the symmetrized direct sum \mathcal{F} = \bigoplus_{N=0}^\infty \mathcal{H}_N, where \mathcal{H}_N = \mathcal{H}_1^{\odot N} denotes the N-particle sector for identical bosons, with the vacuum sector \mathcal{H}_0 = \mathbb{C}. Multi-particle states are generated by applying creation operators a^\dagger(\vec{k}) to the vacuum state |0\rangle, which annihilates all annihilation operators a(\vec{k}) |0\rangle = 0. These operators satisfy the canonical commutation relations [a(\vec{k}), a^\dagger(\vec{k}') ] = (2\pi)^3 2 \omega_k \delta^3(\vec{k} - \vec{k}'), derived from the mode expansion of the field operator \phi(x) = \int \frac{d^3k}{(2\pi)^3 2 \omega_k} \left( a(\vec{k}) e^{-i k \cdot x} + a^\dagger(\vec{k}) e^{i k \cdot x} \right) at equal times. A representative N-particle state in the momentum basis is thus |\vec{k}_1, \dots, \vec{k}_N \rangle = a^\dagger(\vec{k}_1) \cdots a^\dagger(\vec{k}_N) |0\rangle / \sqrt{N!}, which is invariant under Poincaré transformations when the total four-momentum is on-shell. This construction extends to fields with particles and antiparticles by including separate creation operators for each, forming a Fock space that encompasses both sectors in a . However, Haag's theorem reveals fundamental limitations for interacting theories: it proves that no can map the free-field Fock space to an interacting-field while preserving the commutation relations at spacelike separations, implying that interacting fields cannot be defined on the same as free fields in a straightforward manner. This result undermines the standard in perturbative QFT, necessitating alternative approaches like the or algebraic QFT to handle interactions consistently.

Usage in quantum optics and Bose-Einstein condensates

In , Fock space provides the natural framework for describing states of definite photon number in a single electromagnetic mode, denoted as |n\rangle, where n is the number of photons and the vacuum state |0\rangle corresponds to no photons. These Fock states form an orthonormal basis in the for the mode, enabling precise treatments of non-classical light-matter interactions. A seminal application is the Jaynes-Cummings model, which describes a two-level coupled to a quantized field mode via the \hat{H} = \omega a^\dagger a + \frac{\Omega}{2} \sigma_z + g (a \sigma_+ + a^\dagger \sigma_-), where a^\dagger and a are the creation and annihilation operators for photons, \sigma_z, \sigma_+, \sigma_- are atomic Pauli operators, \omega is the field frequency, \Omega the atomic transition frequency, and g the coupling strength. This model, solvable exactly in Fock space, reveals phenomena like Rabi oscillations and collapse-revival dynamics when the field is in a coherent state superposition of Fock states. Optical Fock states exhibit sub-Poissonian photon number statistics, characterized by a variance \Delta n^2 < \langle n \rangle, with the ideal Fock state achieving \Delta n = 0, making them ultimate number-squeezed states useful for quantum and reducing phase noise in . Experimental realizations of such states have been achieved through conditional measurements, such as heralding from parametric down-conversion sources; a 2013 experiment produced multi-photon Fock states up to n=3 in well-defined spatiotemporal modes with fidelities exceeding 80%. These methods project squeezed or coherent states onto Fock components using photon-number-resolving detectors. More recent advances (as of 2024) have demonstrated the of large Fock states up to n=100 photons in superconducting cavities, achieving quantum-enhanced with sensitivity gains of up to 14.8 dB for sensing and approaching the Heisenberg limit. In Bose-Einstein condensates (BECs), Fock space describes the many-body of identical bosons, allowing derivations of effective mean-field descriptions from the full second-quantized . The Gross-Pitaevskii equation, governing the condensate order parameter \psi(\mathbf{r},t), emerges as an approximation for dynamics starting from initial data in Fock space close to the , with rigorous bounds on in the dilute limit where interaction strength scales as $1/N for N particles. This equation, i \hbar \partial_t \psi = [-\frac{\hbar^2}{2m} \nabla^2 + V(\mathbf{r}) + g |\psi|^2] \psi, captures condensate evolution while excitations are treated in the orthogonal Fock subspace. Bogoliubov theory extends this by diagonalizing quadratic fluctuations around the condensate in Fock space, introducing quasiparticle operators \beta_k^\dagger = u_k a_k^\dagger + v_k a_{-k} that create excitations with dispersion E_k = \sqrt{\epsilon_k (\epsilon_k + 2 \mu)}, where \epsilon_k = \hbar^2 k^2 / 2m and \mu is the chemical potential. These Bogoliubov quasiparticles represent collective modes in the BEC, such as sound waves at low momentum, and have been observed in trapped gases through time-of-flight expansion and Bragg spectroscopy.

Mathematical Connections

Relation to Segal-Bargmann space

The Segal–Bargmann space is the consisting of entire analytic functions on \mathbb{C}^d that are square-integrable with respect to the d\mu(z) = \left(1/\pi\right)^d e^{-|z|^2} d^2 z. For the single-mode , the Segal–Bargmann transform defines a unitary map U: L^2(\mathbb{R}) \to \mathcal{HB} from the Schrödinger representation to the Segal–Bargmann space \mathcal{HB}, where \mathcal{HB} denotes the case d=1. This transform is given by the integral (Uf)(z) = \int_{-\infty}^\infty K(x,z) f(x) \, dx, with the Bargmann kernel K(x,z) = \pi^{-1/4} e^{\sqrt{2} z x - z^2/2 - x^2/2}. This construction extends naturally to the bosonic Fock space, which arises as the product over multiple modes or particles, yielding the Segal–Bargmann space on \mathbb{C}^d for d modes. The extension proceeds via coherent states, which serve as an intermediary basis linking the occupation number representation in Fock space to the of holomorphic functions. The transform is an that conjugates the to differential operators on the holomorphic space, satisfying U a U^{-1} = \frac{d}{dz} and U a^\dagger U^{-1} = z. In the multi-mode setting, multiplication by the coordinate z_k acts as the creation operator for the k-th mode, while differentiation \frac{\partial}{\partial z_k} acts as the corresponding annihilation operator.

Fock space and coherent states

Coherent states in Fock space are defined as the eigenstates of the annihilation operator a, satisfying a |\alpha\rangle = \alpha |\alpha\rangle, where \alpha is a eigenvalue. These states can be expressed in the occupation number basis as |\alpha\rangle = e^{-|\alpha|^2/2} \sum_{n=0}^\infty \frac{\alpha^n}{\sqrt{n!}} |n\rangle, or equivalently as the action of the on the : |\alpha\rangle = e^{\alpha a^\dagger - \alpha^* a} |0\rangle. This definition originates from the work of Schrödinger, who constructed such states as minimum uncertainty wave packets for the harmonic oscillator, and was later formalized by Glauber in the context of quantum optics. Key properties of coherent states include their , \langle \alpha | \alpha \rangle = 1, and the fact that they minimize the product for and , achieving the Heisenberg limit \Delta x \Delta p = \hbar/2. The set of coherent states forms an overcomplete basis in Fock space, characterized by the resolution of unity \int \frac{d^2\alpha}{\pi} |\alpha\rangle \langle \alpha | = I, which allows for the expansion of any state in the space as a continuous superposition of coherent states. This overcompleteness property, introduced by Glauber, facilitates applications in and many-body systems by providing a non-orthogonal basis that interpolates between classical and quantum descriptions. For systems with multiple modes, coherent states generalize to |\vec{\alpha}\rangle = \prod_k D(\alpha_k) |0\rangle, where D(\alpha_k) = e^{\alpha_k a_k^\dagger - \alpha_k^* a_k} is the for the k-th , and \vec{\alpha} = (\alpha_1, \alpha_2, \dots) labels the state. The multi-mode resolution of unity follows as \int \prod_k \frac{d^2\alpha_k}{\pi} |\vec{\alpha}\rangle \langle \vec{\alpha} | = I. This extension, also due to Glauber, is essential for describing multi-particle or multi-field configurations in Fock space. Coherent states in Fock space are related to the through a unitary transform that maps them to holomorphic functions in the .

Bargmann-Fock realization of oscillators

The Bargmann-Fock realization offers a holomorphic representation of the within Fock space, mapping states to analytic functions on the . This approach, introduced by Valentine Bargmann, transforms the occupation number basis into a space where quantum operators act via multiplication and differentiation, facilitating calculations in and field theory. In the Fock space occupation number basis, the of the single-mode is given by \hat{H} = \hbar \omega \left( a^\dagger a + \frac{1}{2} \right), where a^\dagger and a are the satisfying [a, a^\dagger] = 1, and the eigenstates |n\rangle (with n = 0, 1, 2, \dots) have eigenvalues (n + 1/2) \hbar \omega. This form arises naturally in the second-quantized description of the oscillator, aligning with the Fock space structure for or modes. In the Bargmann representation, the wave functions corresponding to the number states |n\rangle are expressed as holomorphic functions \psi_n(z) = \frac{z^n}{\sqrt{n!}}, where z \in \mathbb{C} parameterizes the complex plane. The inner product between two states represented by functions \psi(z) and \phi(z) is defined by \langle \psi | \phi \rangle = \int \psi^*(z) \phi(z) \, e^{-|z|^2} \frac{d^2 z}{\pi}, ensuring orthonormality \langle n | m \rangle = \delta_{nm}. This integral is taken over the entire complex plane, with d^2 z = dx \, dy for z = x + i y, and the Gaussian weight enforces the Hilbert space structure. The ladder operators are realized differentially: a \mapsto \frac{d}{dz} and a^\dagger \mapsto z (multiplication by z), reproducing the commutation relations under the given inner product. The position and momentum operators follow from the canonical relations \hat{x} = \sqrt{\frac{\hbar}{2 m \omega}} (a + a^\dagger) and \hat{p} = i \sqrt{\frac{\hbar m \omega}{2}} (a^\dagger - a), yielding realizations \hat{x} \to \sqrt{\frac{\hbar}{2 m \omega}} \left( z + \frac{\partial}{\partial z} \right), \quad \hat{p} \to i \sqrt{\frac{\hbar m \omega}{2}} \left( z - \frac{\partial}{\partial z} \right). These differential operators act on the holomorphic wave functions, preserving the analyticity and enabling exact solutions for oscillator dynamics. For a system of multiple non-interacting oscillators, the Fock space is the of single-mode Fock spaces, realized in the Bargmann-Fock framework as the space of holomorphic functions of several complex variables \mathbf{z} = (z_1, \dots, z_N). The multi-mode wave functions are products \psi_{\mathbf{n}}(\mathbf{z}) = \prod_{k=1}^N \frac{z_k^{n_k}}{\sqrt{n_k!}}, with the inner product extending to \int \psi^*(\mathbf{z}) \phi(\mathbf{z}) e^{-|\mathbf{z}|^2} \frac{d^{2N} \mathbf{z}}{\pi^N}. Operators for each mode act independently on the corresponding variable, such as a_k \mapsto \partial_{z_k} and a_k^\dagger \mapsto z_k, facilitating applications in multi-particle or multi-field systems.

References

  1. [1]
    [PDF] FOCK SPACES - Institut Camille Jordan
    Fock spaces have interpretations in physics, they have been built in order to represent the state space for a system containing an indefinite (and variable).<|control11|><|separator|>
  2. [2]
    [PDF] Fock Space
    Oct 24, 2011 · For fermionic Fock space, each state can either be occupied by one or by no particle. ni ∈ N≥0. (bosons) ni ∈ {0,1} (fermions). We define the ...
  3. [3]
  4. [4]
    [PDF] arXiv:1805.04552v2 [quant-ph] 31 Aug 2018
    Aug 31, 2018 · In particular, the Fock space allows one to build the space of states starting from the single-particle Hilbert space.
  5. [5]
    [PDF] Quantum Theory in Hilbert Space: a Philosophical Review - LSE
    A Hilbert space is a complete inner product space, a complete vector space where every Cauchy sequence converges. It is usually written as H.
  6. [6]
    [PDF] Are Hilbert Spaces Unphysical? Hardly, My Dear! - arXiv
    By way of illustration, for the hydrogen atom with the proton fixed the Hilbert space is L2(R3) with Hamiltonian operator Hhyd, whereas for the harmonic ...
  7. [7]
    [PDF] Impact of relativity on particle localizability and ground state ... - arXiv
    Sep 16, 2019 · The corresponding single-particle Hilbert space (called ... decay sufficiently quickly in momentum space (e.g., Gaussian smearing functions).
  8. [8]
    [PDF] DAMTP - 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. Let's take a moment to understand ...Missing: relativistic | Show results with:relativistic
  9. [9]
    [PDF] A note on the normalization of the momentum eigenfunctions ... - arXiv
    We find that the normalization of these eigenfunctions is a real and not complex number with phase factor chosen equal one (standard books of quantum mechanics) ...
  10. [10]
    3.1: Quantum States of Multi-Particle Systems - Physics LibreTexts
    May 1, 2021 · The symbol ⊗ refers to a tensor product, a mathematical operation that combines two Hilbert spaces to form another Hilbert space. It is most ...
  11. [11]
    [PDF] Chapter 10 Many–particle systems
    In the present chapter we discuss the construction of many particle Hilbert spaces and their application to systems with many identical particles. As ...
  12. [12]
    [PDF] Many-body quantum mechanis
    Consider two particles moving in one dimension. The Hilbert space of each particle is L2(R), the space of wave functions over R. Let {|xi}x∈R be the ...<|control11|><|separator|>
  13. [13]
    Symmetrization Postulate and Its Experimental Foundation
    The symmetrization postulate (SP) states that states of more than one identical particle are either symmetric or antisymmetric under permutations.
  14. [14]
    [PDF] Systems of Identical Particles - CMU Quantum Theory Group
    Mar 21, 2011 · The wave function for a collection of identical bosons must be symmetrical under the interchange of any two particles. Such wave functions lie ...Missing: indistinguishable | Show results with:indistinguishable
  15. [15]
  16. [16]
  17. [17]
    [PDF] arXiv:1405.6570v2 [math.FA] 11 Feb 2015
    Feb 11, 2015 · Definitions and notations. • Let K be a separable Hilbert space. Then the symmetric Fock space Γs(K ) is defined as the direct sum: Γs(K ) ...
  18. [18]
    Konfigurationsraum und zweite Quantelung | Zeitschrift für Physik A ...
    Download PDF · Zeitschrift für Physik. Konfigurationsraum und zweite Quantelung ... Cite this article. Fock, V. Konfigurationsraum und zweite Quantelung. Z ...
  19. [19]
    Zur Quantenmechanik der Gasentartung | Zeitschrift für Physik A ...
    Download PDF · Zeitschrift für Physik. Zur Quantenmechanik der ... Cite this article. Jordan, P. Zur Quantenmechanik der Gasentartung. Z. Physik ...
  20. [20]
    Über das Paulische Äquivalenzverbot | Zeitschrift für Physik A ...
    Download PDF · Zeitschrift für Physik. Über das Paulische Äquivalenzverbot ... Cite this article. Jordan, P., Wigner, E. Über das Paulische Äquivalenzverbot.
  21. [21]
    [PDF] Second Quantization - Arizona Math
    Fock space defined as the direct sum. F(H) := 0 n≥0. L2(Rnν) = C ⊕ L2(Rν) ... On the Fock space the number operator N can be decomposed using operators ...
  22. [22]
    [PDF] Chapter 2 Second Quantisation - TCM
    Second quantisation provides a basic and efficient language in which to formulate many- particle systems. As such, extensive introductions to the concept ...
  23. [23]
    [PDF] Second Quantization
    The Identity Operator 1 of the one-particle Hilbert space. 1 = Xα. |αihα|. (1.156) in Fock space becomes the number operator ˆN. ˆN = Xα â†(α)â(α). (1.157). In ...
  24. [24]
    [PDF] Quantum Theory of Many Particle Systems
    Mar 28, 2023 · Fock space: the creation operator changes the number of particles. ... ing the Fock state ˆρ |n⟩ = n |n⟩ and its conjugate |φ⟩ with ⟨φ ...
  25. [25]
    Introduction to the Hubbard Model
    Mar 9, 2022 · The Fock space repre- sentation is also known as the “second quantization” formalism, but one should note that it is strictly equivalent to the ...
  26. [26]
  27. [27]
  28. [28]
    Fock States - RP Photonics
    Bibliography. [1], V. Fock, “Konfigurationsraum und zweite Quantelung”, Z. Physik 75 (9–10) (1932); doi:10.1007/bf01344458. [2], C. Kurtsiefer, S. Mayer, P ...
  29. [29]
    [PDF] Sub-Poissonian processes in quantum optics - BİLKENT | SCIENCE
    Poissonian photon statistics, and sub-Poissonian states are not necessarily squeezed. ... Fock states can be obtained with photon numbers different from zero.<|control11|><|separator|>
  30. [30]
    [1212.4412] Experimental generation of multi-photon Fock states
    Dec 18, 2012 · Abstract:We experimentally demonstrate the generation of multi-photon Fock states with up to three photons in well-defined spatial-temporal ...
  31. [31]
    [1404.4568] Deriving the Gross-Pitaevskii equation - arXiv
    Apr 17, 2014 · ... Fock space which are energetically close to the ground state and prove that their evolution approximately follows the Gross-Pitaevskii equation.
  32. [32]
    [quant-ph/9912054] Holomorphic Methods in Mathematical Physics
    Dec 11, 1999 · This set of lecture notes gives an introduction to holomorphic function spaces as used in mathematical physics. The emphasis is on the Segal- ...
  33. [33]
  34. [34]
  35. [35]
    Communications on Pure and Applied Mathematics
    Communications on Pure and Applied Mathematics · Volume 14, Issue 3 pp. 187 ... First published: August 1961. https://doi.org/10.1002/cpa.3160140303.
  36. [36]
    On the Representations of the Rotation Group | Rev. Mod. Phys.
    On the Representations of the Rotation Group. V. Bargmann. V. Bargmann Princeton University, Princeton, New Jersey. PDF Share. Rev. Mod. Phys. 34, 829 – ...
  37. [37]
    [PDF] arXiv:math/0110011v1 [math.PR] 1 Oct 2001
    Bargmann, On a Hilbert space of analytic functions and an associated integral transform,. I. Comm. Pure Appl. Math., 14 (1961), 187–214. 10. V. Bargmann, On ...
  38. [38]
    The Bargmann transform and canonical transformations
    May 1, 2002 · V. Bargmann. , “. On a Hilbert space of analytic functions and an associated integral transform. I. ,”. Commun. Pure Appl. Math. 14. ,. 187. –.Missing: original | Show results with:original
  39. [39]
    States of infinitely many oscillators - ScienceDirect
    Infinite quantum systems of harmonic oscillators are investigated in the framework of the Bargmann hilbert space of analytic functions.Missing: original | Show results with:original