Fact-checked by Grok 2 weeks ago

Second quantization

Second quantization is a in for describing systems of many identical particles, employing that act on states in to represent variable particle numbers and inherently incorporate without explicit wave function symmetrization. This approach reformulates the by treating single-particle s as classical fields that are then "quantized" via , distinguishing it from , which relies on symmetrized or antisymmetrized wave functions for fixed particle numbers. The concept emerged in the late 1920s as physicists grappled with many-particle , with introducing key ideas in 1927 through his treatment of radiation fields and particle creation. independently developed the field operator approach around the same time, emphasizing commutation relations for field operators in spacetime, which laid the groundwork for modern . These contributions, building on earlier by , Heisenberg, and , resolved challenges in quantizing interacting fields and identical particles, evolving into a standard tool by the 1930s through works by and others. At its core, second quantization uses as the , spanned by basis states labeled by occupation numbers n_k for each single-particle mode k, allowing superpositions of different particle numbers. For bosons, creation operators a^\dagger_k and annihilation operators a_k satisfy [a_k, a^\dagger_{k'}] = \delta_{kk'}, while for fermions, they obey anticommutation relations \{c_k, c^\dagger_{k'}\} = \delta_{kk'}, enforcing the . Hamiltonians are expressed in terms of these operators, such as the non-interacting form H = \sum_k \epsilon_k a^\dagger_k a_k plus interaction terms like \frac{1}{2} \sum_{k k' q} V_q a^\dagger_{k+q} a^\dagger_{k'-q} a_{k'} a_k for two-body potentials. Field operators \psi^\dagger(\mathbf{r}) and \psi(\mathbf{r}) extend this to position space, enabling descriptions of local densities and currents. In practice, second quantization simplifies calculations in quantum many-body physics, such as deriving the Bose-Einstein condensate order parameter or Fermi liquid properties, and forms the basis for using Green's functions and Feynman diagrams. It is essential for modeling condensed matter phenomena like superconductivity via , electron-phonon interactions, and plasmons in the . Beyond non-relativistic systems, it underpins relativistic , where particles are excitations of underlying fields, facilitating treatments of particle creation and annihilation in high-energy physics.

Fundamentals of Many-Body Quantum Mechanics

First Quantization for Many Particles

In single-particle quantum mechanics, the state of a system is described by a wave function \psi(\mathbf{r}, t) in position space, which evolves according to the time-dependent Schrödinger equation: i \hbar \frac{\partial}{\partial t} \psi(\mathbf{r}, t) = \hat{H} \psi(\mathbf{r}, t), where the Hamiltonian operator \hat{H} = -\frac{\hbar^2}{2m} \nabla^2 + V(\mathbf{r}) incorporates the kinetic energy and an external potential V(\mathbf{r}) for a particle of mass m. This equation provides a complete description of non-relativistic quantum dynamics for isolated particles or those in external fields, allowing computation of observables such as position probabilities via |\psi(\mathbf{r}, t)|^2. The formulation was originally derived by Erwin Schrödinger in 1926 as part of his wave mechanics approach to quantizing atomic systems. For systems of multiple distinguishable particles, the first-quantized formalism extends naturally by constructing a joint wave function as a direct product of individual single-particle wave functions: \psi(\mathbf{r}_1, \mathbf{r}_2, \dots, \mathbf{r}_N; t) = \prod_{i=1}^N \phi_i(\mathbf{r}_i, t). The corresponding is the of the single-particle s, \mathcal{H} = \bigotimes_{i=1}^N \mathcal{H}_i, enabling the total to act separably on each particle's , with interactions added as two-body or higher potentials. This product structure simplifies the treatment of non-interacting or weakly coupled distinguishable systems, such as electrons in different atoms. The construction for multi-particle states was formalized in early by in 1926. When particles are identical, however, the indistinguishability requires the total wave function to transform in a specific manner under particle exchange to avoid overcounting states and ensure physical consistency. According to the symmetrization postulate, the wave function must be either fully symmetric for bosons or fully antisymmetric for fermions under interchange of any two particle coordinates. This requirement arises from the fundamental quantum statistics and enforces phenomena like Bose-Einstein condensation for bosons and the Pauli exclusion principle for fermions. Dirac introduced this postulate in 1926 to reconcile quantum mechanics with the observed behavior of identical particles. For fermionic systems, such as electrons, antisymmetric wave functions are often constructed using Slater determinants, which provide an explicit antisymmetrized product of single-particle orbitals: \psi(\mathbf{r}_1, \dots, \mathbf{r}_N) = \frac{1}{\sqrt{N!}} \det \begin{pmatrix} \phi_1(\mathbf{r}_1) & \phi_1(\mathbf{r}_2) & \cdots & \phi_1(\mathbf{r}_N) \\ \vdots & \vdots & \ddots & \vdots \\ \phi_N(\mathbf{r}_1) & \phi_N(\mathbf{r}_2) & \cdots & \phi_N(\mathbf{r}_N) \end{pmatrix}. This form automatically satisfies antisymmetry and is widely used in approximations like Hartree-Fock theory. The determinant representation was developed by John C. Slater in 1929 to handle complex atomic spectra while respecting fermionic statistics. A concrete example illustrates symmetrization for two identical bosons occupying distinct single-particle states \phi_a(\mathbf{r}) and \phi_b(\mathbf{r}) (with \langle \phi_a | \phi_b \rangle = 0): \psi(\mathbf{r}_1, \mathbf{r}_2) = \frac{1}{\sqrt{2}} \left[ \phi_a(\mathbf{r}_1) \phi_b(\mathbf{r}_2) + \phi_a(\mathbf{r}_2) \phi_b(\mathbf{r}_1) \right]. This normalized symmetric combination ensures the correct exchange symmetry and proper , \int d^3\mathbf{r}_1 d^3\mathbf{r}_2 |\psi|^2 = 1. For fermions, the analogous antisymmetric form would use a minus sign instead of plus, vanishing if \phi_a = \phi_b due to the Pauli principle. Despite its foundational role, the first-quantized approach faces significant challenges for many-particle systems. The configuration space dimensionality scales exponentially with the particle number N, as d^{3N} for d spatial basis functions per particle in three dimensions, rendering exact numerical solutions computationally infeasible beyond small N (typically N \lesssim 10) even on modern hardware. This "curse of dimensionality" arises directly from the high-dimensional required to capture correlations. Additionally, incorporating via symmetrization or antisymmetrization complicates the basis representation and increases computational overhead, while strong interactions demand variational methods or approximations that struggle to scale efficiently. These limitations become acute in condensed matter and , where N can reach $10^{23}. offers an alternative representation that mitigates some of these issues by focusing on occupation numbers rather than explicit coordinates.

Motivation for Second Quantization

In , the description of many-particle systems relies on wave functions in a of fixed dimension, which becomes increasingly cumbersome as the number of particles grows large. For , the wave function must be explicitly symmetrized (for bosons) or antisymmetrized (for fermions) to enforce the correct exchange statistics, a process that involves constructing Slater determinants or permanents and leads to inefficient computations for systems with thousands or more particles, such as those in . Moreover, assumes a conserved particle number, rendering it inadequate for scenarios involving particle creation or annihilation, such as strong interactions or relativistic processes where occurs. Second quantization addresses these limitations by reformulating the in terms of an of acting on a , where particles are treated as excitations of an underlying field rather than fundamental entities with fixed coordinates. This approach naturally incorporates indistinguishability through commutation relations—bosonic for [a, a†] = 1 or fermionic for {a, a†} = 1—eliminating the need for manual symmetrization and enabling efficient handling of interactions via operator expansions. It also accommodates variable particle numbers by constructing the as a over sectors of different particle counts, making it ideal for grand canonical ensembles prevalent in and condensed matter systems. Historically, the motivation arose in the late 1920s from the need to reconcile quantum mechanics with relativity and to describe radiation as quantized fields, where photons could be emitted or absorbed, implying non-conserved particle numbers. Paul Dirac introduced the core idea in 1927 by quantizing the electromagnetic field to explain emission and absorption, treating field amplitudes as operators that create or destroy quanta. Pascual Jordan and Eugene Wigner extended this in 1928 to fermionic systems, providing a systematic operator formalism for identical particles. This framework proved essential for quantum electrodynamics and later quantum field theories, where first quantization fails due to issues like negative probabilities in relativistic single-particle equations. Conceptually, second quantization maps first-quantized s to s without altering the underlying physics; for instance, a single-particle ψ(x) becomes a ˆψ(x) whose values recover the original probabilities, but now acting on a larger that includes multi-particle states. This shifts focus from explicit s to algebraic manipulations, facilitating and diagrammatic techniques for complex interactions. The key insight is that the total many-body emerges as the ⊕_N ℋ_N over particle number sectors ℋ_N, allowing seamless transitions between sectors via s and providing a unified description for both non-relativistic many-body problems and relativistic theories.

Fock Space and Second-Quantized States

Construction of Fock Space

In second quantization, the serves as the for systems of identical particles where the total number of particles is not fixed, allowing for superpositions across different particle numbers. It is constructed as the of the N-particle s for all possible occupation numbers N, mathematically expressed as \mathcal{F} = \bigoplus_{N=0}^{\infty} \mathcal{H}_N, where \mathcal{H}_N denotes the of N , and the N=0 sector is the trivial \mathbb{C}. For identical particles obeying Bose-Einstein statistics, each \mathcal{H}_N is the symmetric tensor product of the single-particle \mathcal{H}_1, ensuring that the multi-particle wave functions remain invariant under particle exchange. In contrast, for particles following Fermi-Dirac statistics, \mathcal{H}_N is the antisymmetric tensor product (or , using products), which enforces antisymmetry and the by changing sign under odd permutations. These symmetry requirements directly incorporate the indistinguishability of particles into the structure of the space, avoiding overcounting in the construction. The vacuum state |0\rangle, or empty state, occupies the N=0 sector and acts as the reference from which all other states are built, representing the absence of particles with \langle 0 | 0 \rangle = 1. To illustrate, consider a single-particle \mathcal{H}_1 spanned by an \{|k\rangle\}_k, such as or eigenstates; then, states in \mathcal{H}_N for bosons consist of symmetric sums over permutations of these basis vectors, normalized appropriately, while fermionic states use antisymmetric combinations like Slater determinants for distinct occupations.

Fock States for Identical Particles

In the second quantization formalism for identical bosons, the basis states of are specified by occupation numbers | n_1, n_2, \dots \rangle, where n_i \geq 0 represents the arbitrary non-negative number of bosons in the single-particle labeled by i. These states account for the indistinguishability and Bose-Einstein of the particles, allowing multiple bosons to occupy the same without restriction. The corresponding wave function in the first-quantized is a fully symmetrized product of single-particle wave functions, constructed as a permanent to ensure under particle . For instance, in a two- system, the state |2,1\rangle describes three bosons with two in mode 1 (characterized by wave function \phi_1(\mathbf{x})) and one in mode 2 (\phi_2(\mathbf{x})); its position-space is the permanent \Psi(\mathbf{x}_1, \mathbf{x}_2, \mathbf{x}_3) = \frac{1}{\sqrt{6}} \left[ \phi_1(\mathbf{x}_1) \phi_1(\mathbf{x}_2) \phi_2(\mathbf{x}_3) + \phi_1(\mathbf{x}_1) \phi_2(\mathbf{x}_2) \phi_1(\mathbf{x}_3) + \phi_2(\mathbf{x}_1) \phi_1(\mathbf{x}_2) \phi_1(\mathbf{x}_3) \right], normalized such that \int |\Psi|^2 d\mathbf{x}_1 d\mathbf{x}_2 d\mathbf{x}_3 = 1. The bosonic Fock states form an , with \langle n_1, n_2, \dots | m_1, m_2, \dots \rangle = \delta_{n_1 m_1} \delta_{n_2 m_2} \cdots, ensuring mutual for distinct occupation configurations. For identical fermions, the Fock states are similarly labeled by occupation numbers but restricted by the , so | \{ n_k \} \rangle with each n_k = 0 or $1, indicating whether mode k is empty or singly occupied. The curly braces emphasize the set notation, as the of modes is immaterial due to the antisymmetric nature of fermionic states. The first-quantized is an antisymmetrized product, given by a to enforce antisymmetry. For a two-mode fermionic state with one fermion in each mode, | \{1,1\} \rangle, the is \Psi(\mathbf{x}_1, \mathbf{x}_2) = \frac{1}{\sqrt{2}} \det \begin{pmatrix} \phi_1(\mathbf{x}_1) & \phi_2(\mathbf{x}_1) \\ \phi_1(\mathbf{x}_2) & \phi_2(\mathbf{x}_2) \end{pmatrix} = \frac{1}{\sqrt{2}} \left[ \phi_1(\mathbf{x}_1) \phi_2(\mathbf{x}_2) - \phi_1(\mathbf{x}_2) \phi_2(\mathbf{x}_1) \right], which vanishes if both fermions attempt to occupy the same mode, upholding the exclusion principle. Like their bosonic counterparts, fermionic Fock states are normalized and orthogonal within the basis, \langle \{ n_k \} | \{ m_l \} \rangle = \delta_{\{n\},\{m\}}.

Creation and Annihilation Operators

General Properties and Insertion/Deletion

In second quantization, serve as fundamental tools for constructing and manipulating states in , which encompasses sectors with varying particle numbers. The creation operator a^\dagger_k adds a single particle in the single-particle state labeled by k to an N-particle state, mapping it to an (N+1)-particle state while preserving the required under particle exchange. Conversely, the annihilation operator a_k removes a particle from the state k in an N-particle state, mapping it to an (N-1)-particle state, again maintaining the appropriate statistics. These operators act linearly on the Fock states that span the space. The process of insertion via a^\dagger_k incorporates the new particle in a manner consistent with the identical particle statistics, ensuring the resulting state remains properly symmetrized or antisymmetrized. Deletion through a_k similarly extracts the particle without violating these statistical constraints, projecting the state onto the lower-particle-number sector. This framework allows for a unified description of systems where particle number is not fixed, facilitating the treatment of phenomena like particle creation and absorption. The creation and annihilation operators are related by Hermitian conjugation, satisfying the adjoint relation (a_k)^\dagger = a^\dagger_k. This ensures that the operators preserve the inner product structure of the , with matrix elements transforming appropriately under conjugation. The |0\rangle, representing the zero-particle sector of , is annihilated by all : a_k |0\rangle = 0 for any k. Applying a creation to the vacuum generates a one-particle state, up to : a^\dagger_k |0\rangle \propto |k\rangle, and the vacuum remains invariant under further creations in the sense that it serves as the foundational state for building higher occupancy. In general, these operators satisfy algebraic relations that depend on the particle statistics: for bosons, the commutator [a_k, a^\dagger_l] = \delta_{kl}, while for fermions, the anticommutator \{a_k, a^\dagger_l\} = \delta_{kl}. These relations underpin the distinct behaviors in subsequent specializations but are introduced here as the universal structure enabling particle number changes.

Bosonic Operators

In second quantization, bosonic creation and annihilation operators a_k and a_k^\dagger for mode k are defined to satisfy the commutation relations [a_k, a_l^\dagger] = \delta_{kl}, [a_k, a_l] = 0, and [a_k^\dagger, a_l^\dagger] = 0. These relations generalize the of the to an infinite set of modes, enabling the description of systems with indistinguishable bosons where multiple particles can occupy the same state. This formalism was first introduced by Dirac in his treatment of the quantized , where the operators correspond to the absorption and emission of photons. The action of these operators on Fock states, which form the basis of the bosonic , is given by a_k^\dagger |\{n\}\rangle = \sqrt{n_k + 1} |\{n + \mathbf{e}_k\}\rangle for and a_k |\{n\}\rangle = \sqrt{n_k} |\{n - \mathbf{e}_k\}\rangle for , where |\{n\}\rangle denotes a state with occupation numbers n_k in mode k, and \mathbf{e}_k is the unit vector in the k-th direction. The square-root factors ensure unitarity and consistency with the commutation relations, allowing arbitrary occupation numbers without Pauli exclusion. This explicit action was formalized by Fock in his construction of the configuration space for second quantization. The number for mode k is defined as N_k = a_k^\dagger a_k, which counts the particles in that and satisfies [N_k, a_l^\dagger] = \delta_{kl} a_l^\dagger (with similar relations for a_l). These commutators follow directly from the bosonic and facilitate the of values in many-body states. In the abstract basis, the operators act independently on each , mirroring the non-interacting oscillators. The Baker-Campbell-Hausdorff applies to products of exponentials involving bosonic operators, simplifying expressions like the D(\alpha) = \exp(\alpha a_k^\dagger - \alpha^* a_k), which generates coherent states from the . This relation, \exp(A) \exp(B) = \exp(A + B + \frac{1}{2}[A, B]) for [A, [A, B]] = [B, [A, B]] = 0, is crucial for deriving the of displaced operators, such as D^\dagger(\alpha) a_k D(\alpha) = a_k + \alpha. The plays a key role in for describing fields. As an example, in the position representation for a single mode, the annihilation operator realizes the harmonic oscillator lowering operator a = \sqrt{\frac{m \omega}{2 \hbar}} \left( x + \frac{i p}{m \omega} \right), with a^\dagger its adjoint, where x and p are the position and momentum operators satisfying [x, p] = i \hbar. This concrete form connects the abstract second-quantized modes to first-quantized single-particle mechanics, extended to many particles via the Fock basis.

Fermionic Operators

Fermionic creation and annihilation operators, denoted a_k and a_k^\dagger respectively, where k labels single-particle states, are defined to satisfy the canonical anticommutation relations (CAR): \{a_k, a_l^\dagger\} = \delta_{kl}, \quad \{a_k, a_l\} = 0, \quad \{a_k^\dagger, a_l^\dagger\} = 0. These relations ensure the antisymmetric nature of fermionic wave functions and enforce the Pauli exclusion principle. The action of these operators on fermionic Fock states |{n}\rangle, where \{n\} specifies the occupation numbers n_k = 0 or $1 for each mode k, is given by a_k^\dagger |\{n\}\rangle = \sqrt{1 - n_k} |\{n + e_k\}\rangle, a_k |\{n}\}\rangle = \sqrt{n_k} |\{n - e_k\}\rangle, with e_k the unit vector in the k-th direction. This reflects the inability to create a second fermion in an occupied state (n_k = 1) or annihilate from an empty one (n_k = 0). The number operator for mode k is N_k = a_k^\dagger a_k, which is idempotent: N_k^2 = N_k. Its eigenvalues are 0 or 1, projecting onto occupied or unoccupied states, and satisfying N_k (1 - N_k) = 0. In lattice models, such as one-dimensional tight-binding systems, fermionic operators are often mapped to spin operators via the , which introduces non-local operators to preserve the anticommutation relations across sites. For spins at sites j < l, the transformation involves a_l = \left( \prod_{m=1}^{l-1} \sigma_m^z \right) \sigma_l^-, where \sigma are Pauli operators, ensuring locality in the limit but introducing correlations in higher dimensions. A simple example is a single-site (single-mode) : a^\dagger |0\rangle = |1\rangle, a |1\rangle = |0\rangle, and a |0\rangle = 0, illustrating the binary occupation enforced by the CAR. These operators adapt the general concept of particle insertion and deletion in second quantization to fermionic statistics through the use of anticommutators, guaranteeing antisymmetric multi-particle states.

Quantum Field Operators

Definition in Position Representation

In second quantization, the quantum field operators in position representation, denoted as \hat{\psi}(\mathbf{r}) and \hat{\psi}^\dagger(\mathbf{r}), serve as the continuous-space analogs of the discrete , enabling the description of particle creation and annihilation at specific positions \mathbf{r} in a many-body . These operators act on the , where \hat{\psi}^\dagger(\mathbf{r}) creates a particle at position \mathbf{r}, and \hat{\psi}(\mathbf{r}) annihilates one there, facilitating a basis-independent formulation for systems with variable particle numbers. The field operators are expanded in terms of a complete of single-particle orbitals \{\phi_\alpha(\mathbf{r})\}, where \alpha labels the basis states (e.g., or eigenstates including ). Specifically, \hat{\psi}(\mathbf{r}) = \sum_\alpha \phi_\alpha(\mathbf{r}) \hat{a}_\alpha, \quad \hat{\psi}^\dagger(\mathbf{r}) = \sum_\alpha \phi_\alpha^*(\mathbf{r}) \hat{a}_\alpha^\dagger, with \hat{a}_\alpha and \hat{a}_\alpha^\dagger being the annihilation and creation operators for the discrete modes \alpha. This expansion bridges the discrete mode description to the continuous space, where the orbitals \phi_\alpha(\mathbf{r}) satisfy \int d\mathbf{r} \, \phi_\alpha^*(\mathbf{r}) \phi_\beta(\mathbf{r}) = \delta_{\alpha\beta}. The algebraic properties of these field operators depend on the particle statistics. For bosons, they satisfy the commutation relations [\hat{\psi}(\mathbf{r}), \hat{\psi}^\dagger(\mathbf{r}')] = \delta(\mathbf{r} - \mathbf{r}'), \quad [\hat{\psi}(\mathbf{r}), \hat{\psi}(\mathbf{r}')] = 0, \quad [\hat{\psi}^\dagger(\mathbf{r}), \hat{\psi}^\dagger(\mathbf{r}')] = 0, while for fermions, the anticommutation relations hold: \{\hat{\psi}(\mathbf{r}), \hat{\psi}^\dagger(\mathbf{r}')\} = \delta(\mathbf{r} - \mathbf{r}'), \quad \{\hat{\psi}(\mathbf{r}), \hat{\psi}(\mathbf{r}')\} = 0, \quad \{\hat{\psi}^\dagger(\mathbf{r}), \hat{\psi}^\dagger(\mathbf{r}')\} = 0. These relations ensure the correct symmetry of the many-body wave function under particle exchange and follow from the corresponding of the discrete operators in the limit of a complete basis. The field operators are instrumental in constructing local observables, such as the operator \hat{\rho}(\mathbf{r}) = \hat{\psi}^\dagger(\mathbf{r}) \hat{\psi}(\mathbf{r}), which represents the particle density at \mathbf{r} and integrates to the total particle number operator \hat{N} = \int d\mathbf{r} \, \hat{\rho}(\mathbf{r}). Similar bilinears yield operators, providing a natural framework for describing spatially resolved quantities in interacting many-body systems. This position-space formulation arises as the continuum limit of the -mode expansion, where a dense basis of modes (e.g., plane waves in a large volume) leads to the of sums over indices by integrals, with the [Dirac delta function](/page/Dirac delta function) emerging from the \sum_\alpha \phi_\alpha^*(\mathbf{r}) \phi_\alpha(\mathbf{r}') = \delta(\mathbf{r} - \mathbf{r}').

Connection to Particle Operators

In second quantization, the quantum field operators bridge the continuous position representation with the discrete mode or particle basis through a Fourier transform relation, particularly evident in momentum space for plane wave modes. For a system of non-interacting particles in a volume V, the field annihilation operator in position space is expressed as \hat{\psi}(\mathbf{r}) = \frac{1}{\sqrt{V}} \sum_{\mathbf{k}} e^{i \mathbf{k} \cdot \mathbf{r}} \hat{a}_{\mathbf{k}}, where \hat{a}_{\mathbf{k}} annihilates a particle in the momentum mode \mathbf{k}, and the sum is over discrete wavevectors in a box normalization. The inverse relation is \hat{a}_{\mathbf{k}} = \frac{1}{\sqrt{V}} \int d^3 r \, e^{-i \mathbf{k} \cdot \mathbf{r}} \hat{\psi}(\mathbf{r}), demonstrating that the field operator is a superposition of particle annihilation operators weighted by plane wave functions, with the commutation relations [\hat{a}_{\mathbf{k}}, \hat{a}^\dagger_{\mathbf{k}'}] = \delta_{\mathbf{k},\mathbf{k}'} (for bosons) preserved under this unitary Fourier transformation. When acting on Fock states, which are eigenstates of particle number in the mode basis labeled by numbers \{n_{\mathbf{k}}\}, the field creation operator \hat{\psi}^\dagger(\mathbf{r}) generates a state with a particle delocalized across modes but effectively localized at \mathbf{r}. Specifically, \hat{\psi}^\dagger(\mathbf{r}) |\{n_{\mathbf{k}}\}\rangle creates a one-particle superposed over momentum s, equivalent to \sum_{\mathbf{k}} e^{-i \mathbf{k} \cdot \mathbf{r}} / \sqrt{V} \, \hat{a}^\dagger_{\mathbf{k}} |\{n_{\mathbf{k}}\}\rangle, where the encodes the . This action connects the second-quantized description to first-quantized wave mechanics, as the resulting state has a proportional to the superposition. For fermions, the antisymmetric nature ensures the Pauli exclusion, with \hat{\psi}^\dagger(\mathbf{r}) unable to create a particle if the local is already maximal. The equivalence between field and particle operators is further highlighted by expectation values, which recover first-quantized observables. The particle \hat{n}(\mathbf{r}) = \hat{\psi}^\dagger(\mathbf{r}) \hat{\psi}(\mathbf{r}) has an expectation value \langle \{n_{\mathbf{k}}\} | \hat{n}(\mathbf{r}) | \{n_{\mathbf{k}}\} \rangle = \sum_{\mathbf{k}} n_{\mathbf{k}} / V for uniform plane waves, matching the probability from the first-quantized single-particle wave function |\phi(\mathbf{r})|^2. More generally, for a mixed state, \langle \hat{\psi}^\dagger(\mathbf{r}) \hat{\psi}(\mathbf{r}') \rangle yields the one-body , directly analogous to the overlap of first-quantized orbitals. Under unitary basis transformations, such as rotations or boosts, both field and particle s transform covariantly: \hat{a}^\dagger_{\tilde{\mathbf{k}}} = \sum_{\mathbf{k}} U_{\mathbf{k},\tilde{\mathbf{k}}} \hat{a}^\dagger_{\mathbf{k}}, ensuring the commutation relations remain invariant and preserving physical observables like total particle number \hat{N} = \sum_{\mathbf{k}} \hat{a}^\dagger_{\mathbf{k}} \hat{a}_{\mathbf{k}} = \int d^3 r \, \hat{n}(\mathbf{r}). A concrete example is the free particle in a one-dimensional box of length L, where modes are \mathbf{k} = 2\pi n / L for integer n. The field operator becomes \hat{\psi}(x) = \frac{1}{\sqrt{L}} \sum_n e^{i k_n x} \hat{a}_{k_n}, and the kinetic energy Hamiltonian is \hat{H} = \sum_n \frac{\hbar^2 k_n^2}{2m} \hat{a}^\dagger_{k_n} \hat{a}_{k_n} = \int_0^L dx \, \hat{\psi}^\dagger(x) \left( -\frac{\hbar^2}{2m} \frac{d^2}{dx^2} \right) \hat{\psi}(x), illustrating the exact mapping between particle-mode and position-space forms while maintaining the discrete spectrum of the box. This setup underscores how second quantization unifies delocalized particle excitations with local field descriptions.

Historical Context and Applications

Key Developments and Contributors

The development of second quantization emerged in the late 1920s as physicists sought to reconcile with the behavior of identical particles and radiation fields. Early ideas drew analogies between quantum mechanical observables and field descriptions, building on Werner Heisenberg's 1925 formulation of , which laid the groundwork for treating dynamical systems in terms of spectral frequencies akin to field modes. This approach was extended in 1926 by , Heisenberg, and , who applied quantization procedures to the in their seminal paper, marking an initial step toward a of fields. A pivotal advancement came in 1927 with Paul Dirac's work on the quantum theory of radiation emission and absorption, where he introduced operator methods to describe particle interactions with quantized fields, effectively bridging single-particle and field theory. This laid the foundation for handling variable particle numbers. In 1928, and further developed these ideas for fermions, introducing anticommutation relations to enforce the in a many-body context. Vladimir Fock's 1932 contribution formalized the mathematical structure by defining the configuration space for systems with indefinite particle numbers, now known as , which provided a rigorous framework for second quantization. During , , along with Heisenberg, refined these concepts in the context of (), addressing infinities and interactions in their 1929–1930 papers on the quantum dynamics of wave fields. Post-World War II, second quantization became integral to many-body theory and renormalized QED. Julian Schwinger's 1948 covariant formulation and Richard Feynman's 1949 space-time approach revitalized , incorporating second quantization to handle divergent integrals and particle creation/annihilation processes effectively. By the , these methods had evolved into a standard tool for and particle theory, enabling systematic treatments of interacting systems.

Modern Applications

In , second quantization provides a powerful framework for describing strongly correlated electron systems, particularly through the , which captures the competition between and on-site repulsion. The model's , expressed using fermionic , enables precise modeling of phenomena such as Mott insulation and in materials like cuprates. This formalism has been instrumental in numerical methods like exact diagonalization and , revealing phase transitions driven by electron correlations. Similarly, in the Bardeen-Cooper-Schrieffer ( of conventional , second quantization reformulates the electron-phonon interaction into a pairing Hamiltonian, where operators create and annihilate Cooper pairs of opposite spin and momentum. This approach explains the formation of a superconducting condensate and the exponential temperature dependence of the energy gap, underpinning applications in materials like niobium-based superconductors. The second-quantized BCS , a coherent superposition of paired states, has been extended to projected BCS variants for handling number conservation in finite systems. In , second quantization describes electromagnetic fields in terms of photon creation and annihilation , facilitating the study of coherent states—minimum-uncertainty Gaussian wavepackets that model classical light while incorporating quantum fluctuations. These states are central to (), where they characterize photon statistics in laser-driven cavities interacting with atoms, enabling phenomena like Rabi oscillations and . Seminal work established coherent states as eigenstates of the annihilation , bridging classical and quantum descriptions in experiments with optical cavities. Quantum information science leverages second quantization to model fermionic systems hosting Majorana zero modes, self-conjugate particles that enable topological s with inherent protection against local noise. In semiconductor-superconductor hybrids, Majorana operators, satisfying anticommutation relations akin to standard fermions but with reality constraints, encode qubits non-locally across wire ends, supporting braiding operations for fault-tolerant computation. This approach has driven proposals for scalable quantum processors, with experimental signatures observed in setups; as of 2025, recent progress includes Microsoft's announcement of a topological qubit device in February 2025, though it faced community scrutiny, and other experiments demonstrating multiple Majorana modes in 2024. In nuclear physics, the shell model uses second quantization to construct many-body wavefunctions from single-nucleon creation operators acting on a fermionic vacuum, accounting for Pauli exclusion and residual interactions. This method reproduces energy levels and electromagnetic transitions in medium-mass nuclei, such as sd-shell isotopes, by diagonalizing effective Hamiltonians in truncated basis spaces. Modern extensions, like no-core shell model calculations, apply the formalism ab initio from nucleon-nucleon potentials, improving predictions for light nuclei spectra. Although second quantization originated in non-relativistic many-body theory, its relativistic extension underpins for particle creation in high-energy processes; however, it retains a non-relativistic focus in ultracold atomic gases, where field operators describe Bose-Einstein condensates and Fermi seas in optical lattices. In these systems, the Gross-Pitaevskii equation emerges from the second-quantized for weakly interacting bosons, simulating Hubbard-like models for quantum phase transitions. This has enabled analog quantum simulations of condensed matter phenomena using tunable atom traps.

Nomenclature and Clarifications

Origin of the Term

The term "second quantization" was coined by the Soviet physicist in his 1932 paper titled Konfigurationsraum und zweite Quantelung, where he introduced the configuration space formalism for many-particle quantum systems. This built directly on the pioneering work of in the late , who developed the operator methods for quantizing fields and handling identical particles, as seen in his 1927 collaboration with on the quantization of radiation fields. Jordan's contributions, including the introduction of for bosonic systems, laid the groundwork for treating quantum fields as operators acting on a of variable particle number. The nomenclature distinguishes "," which refers to the standard quantization of single-particle mechanics (e.g., promoting classical and to operators in Schrödinger's ), from "," which extends this to fields or ensembles where particle number is not fixed, allowing for and processes. However, the term is widely regarded as a , as it does not represent a genuine second stage of quantization but rather a unified reformulation of suitable for and interacting systems, avoiding the limitations of fixed-particle-number wave functions. This perspective emphasizes that second quantization is essentially the language of adapted to many-body problems, rather than an additional layer atop . Alternative designations, such as "occupation number representation" (highlighting the basis of states labeled by particle occupation in single-particle modes) or "many-body formalism," better capture its role without implying a sequential quantization process. These terms underscore the focus on as the natural arena for describing systems with indefinite particle counts. Historically, the framework originated in the context of relativistic during the early 1930s, where it was essential for quantizing fields like the while respecting and . Over time, it was generalized to non-relativistic for condensed matter and applications, such as electron gases and Bose-Einstein condensates, proving indispensable for handling interactions in systems of identical particles.

Common Misconceptions

A common misconception about second quantization is that it is applicable only to relativistic (QFT), where particle and are prominent. In fact, second quantization is indispensable in non-relativistic many-body physics, providing a systematic way to describe systems of identical particles, such as electrons in solids or ultracold atoms, by constructing the as a and using to handle indistinguishability and interactions efficiently. This simplifies calculations for fixed or variable particle numbers without invoking relativistic effects, as demonstrated in treatments of Bose-Einstein condensates and fermionic gases. Scholarly reviews emphasize its broad utility beyond , underscoring that the technique originated in non-relativistic contexts before its extension to QFT. Another frequent confusion arises from the terminology, leading some to believe that second quantization involves "quantizing the wave function" produced by , implying a double application of quantization rules. Instead, it treats the wave function (or single-particle orbital) as a classical that is promoted to an acting on the many-body , thereby quantizing the underlying configuration space rather than re-quantizing an already quantum object. This distinction clarifies that the process constructs operators like \hat{\psi}(\mathbf{r}) whose eigenvalues relate to particle densities, avoiding any redundant quantization of the Schrödinger wave function itself. It is not a literal second quantization but a reformulation to accommodate many particles, as highlighted in foundational expositions. Second quantization is sometimes conflated with other formalisms like path integrals or density matrices, but these serve distinct purposes. Path integrals provide a functional representation of quantum evolution, summing over configurations, whereas second quantization relies on in for direct computation of expectation values and dynamics. Density matrices, useful for mixed states and reduced descriptions in many-body systems, do not inherently incorporate and processes; second quantization uniquely facilitates this through its operator structure, though both can describe similar statistical properties. This operator-based approach complements but does not replace these methods, offering advantages in perturbative expansions for interacting systems. In the context of lattice models, the Jordan-Wigner transformation—essential for mapping fermionic Hamiltonians to operators—introduces non-local operators to enforce anticommutation relations, which is occasionally misinterpreted as a fundamental flaw or artifact. Far from a defect, this non-locality is a necessary feature arising from the topological requirements of representing fermionic statistics on a one-dimensional , ensuring the correct without altering the physical locality of interactions. Reviews of the transformation affirm its exactness in solvable models like the XY chain, where the non-locality enables analytic solutions while preserving locality in the original fermionic description. Finally, second quantization is often viewed as introducing a novel physical theory, but it is precisely a representational tool: an equivalent reformulation of many-particle using field operators and , which generalizes the fixed-number of to include superselection sectors for varying particle numbers. This equivalence holds for non-interacting cases, with interactions added via operator products, and it does not alter the underlying quantum principles but enhances computational tractability. As a mathematical framework, it bridges non-relativistic and relativistic regimes without claiming new physics.

References

  1. [1]
    [PDF] Second Quantization of Identical Bosons - UT Physics
    Second quantization uses creation/annihilation operators in Fock space, not a new quantization, but a re-written quantum mechanics. It's used for identical ...
  2. [2]
    None
    Below is a merged summary of the "Second Quantization" sections from the book *Many-Body Quantum Theory in Condensed Matter Physics* by Henrik Bruus and Karsten Flensberg. The information is consolidated into a dense, structured format using a table to retain all details efficiently, followed by a narrative summary that integrates the key points. The table captures the core elements (Definition, Comparison, Key Formalism, Applications) across the various chapters, while the narrative provides a cohesive overview.
  3. [3]
    None
    ### Summary of Second Quantization in arXiv:1501.07384
  4. [4]
    [PDF] Physics 212C: Advanced Quantum Mechanics Spring 2020
    Dec 8, 2020 · Relativistic quantum mechanics. Techniques: How to describe an indefinite number of particles ('second quantization'). Hartree-Fock, Thomas- ...
  5. [5]
    [PDF] Quantum Field Theory - DAMTP
    The requirement of locality remains a strong motivation for studying field theories in the quantum world. ... 5For more details on the history of quantum field ...<|control11|><|separator|>
  6. [6]
    Second quantization - Scholarpedia
    Jun 28, 2010 · History. Second quantization was introduced by Dirac as an algorithm for the construction of Quantum Mechanics of assemblies of identical ...History · The Fock space for Bosons · Quantization of the Free Bose...Missing: motivation | Show results with:motivation
  7. [7]
  8. [8]
    [PDF] Second Quantization - Arizona Math
    For the Fock space corresponded to the Hilbert space H = L2(Rnν), there are two type of restrictions imposed by quantum statistics. These restrictions are.
  9. [9]
    [PDF] Chapter 3: Elements of second quantization
    Somewhat pedantically, we will introduce the Fock space: This separable Hilbert space is the direct sum of Hilbert spaces of fixed particle number, that is.<|control11|><|separator|>
  10. [10]
    [PDF] Quantum Theory of Many Particle Systems
    Mar 28, 2023 · A. L. Fetter and J. D. Walecka, Quantum Theory of Many-Particle Systems (McGraw Hill, ... creation and annihilation oper- ators: Number ...
  11. [11]
    [PDF] Many-body quantum theory in condensed matter physics
    This introduction to quantum field theory in condensed matter physics has emerged from our courses for graduate and advanced undergraduate students at the Niels ...
  12. [12]
  13. [13]
    The quantum theory of the emission and absorption of radiation
    The new quantum theory, based on the assumption that the dynamical variables do not obey the commutative law of multiplication, has by now been developed ...
  14. [14]
    [PDF] P. Jordan, E. Wigner. Über das Paulischen Äquivalenzverbot
    (5). Über das Paulische Äquivalenzverbot. Von P. Jordan und E. Wigner in Göttingen. (Eingegangen am 26. Januar 1928.) Die Arbeit enthält eine Fortsetzung der ...
  15. [15]
    [PDF] Fermionic Algebra and Fock Space - UT Physics
    ... bosonic and the fermionic Fock spaces have similar forms, the actual operators are quite different because the two Fock spaces have different algebras of ...
  16. [16]
    [PDF] Many-body systems - UBC Physics
    In the second quantization, it is as if we now also quantize the wavefunction into the field operator ˆΨ(r). Now, let us simplify the notation and remove the ...
  17. [17]
    [PDF] Chapter 2 Second Quantisation - TCM
    the Fock space in general and any N-particle Hilbert space in particular, it must be possible to represent any operator ˆO1 in an a-representation. Now, ...
  18. [18]
    [PDF] Introduction to Second Quantization Contents 1 Preliminaries
    Second quantization is an efficient tool describing many-body systems compactly, using different state labeling and creation/annihilation operators.
  19. [19]
    [PDF] Second Quantization
    In summary, after quantizing the elastic solid we found that the quantum states of this system can be classified in terms if a set of excitations, the longi-.
  20. [20]
    [PDF] Chap 1 Manybody wave function and Second quantization
    In the formulation of second quantization, operators are written in creation and anni- hilation operators. A. Occupation number representation. After ...
  21. [21]
    Zur Quantenmechanik. II. | Zeitschrift für Physik A Hadrons and nuclei
    Download PDF · Zeitschrift für Physik. Zur Quantenmechanik. II. Download ... Cite this article. Born, M., Heisenberg, W. & Jordan, P. Zur Quantenmechanik. II..
  22. [22]
    Konfigurationsraum und zweite Quantelung | Zeitschrift für Physik A ...
    Fock, V. Konfigurationsraum und zweite Quantelung. Z. Physik 75, 622–647 (1932). https://doi.org/10.1007/BF01344458. Download citation. Received: 10 March 1932.
  23. [23]
    Space-Time Approach to Quantum Electrodynamics | Phys. Rev.
    R. P. Feynman, Phys. Rev. 76, 749 (1949); R. P. Feynman, Phys. Rev. 74, 939 (1948); R. P. Feynman, Phys. Rev. 74, 1430 (1948); J. Schwinger, Phys. Rev. 74, 1439 ...
  24. [24]
    [PDF] The Hubbard Model arXiv:2103.12097v2 [cond-mat.str-el] 18 Jul 2021
    Jul 18, 2021 · The repulsive Hubbard model has been immensely useful in understand- ing strongly correlated electron systems, and serves as the ...
  25. [25]
    From second quantization to the BCS Hamiltonian - ScienceDirect.com
    Under the conditions k ′ + k =0 and σ ′ + σ =0 one can straightforwardly derive the Bardeen–Cooper–Schrieffer (BCS) Hamiltonian from the second quantization ...Missing: formulation paper
  26. [26]
    [PDF] Introduction to topological superconductivity and Majorana fermions
    Nov 16, 2012 · The text should be accessible for readers with a basic understanding of quantum mechanics and second quantization, and does not require ...
  27. [27]
    [PDF] Recent developments in no-core shell-model calculations - arXiv
    Apr 2, 2009 · The powerful techniques based on the second quantization and developed for standard shell model calculations can then be utilized. Therefore, ...<|separator|>
  28. [28]
    [PDF] Tools for quantum simulation with ultracold atoms in optical lattices
    This Technical Review provides an accessible source of technical references especially targeted at newcomers to the field of experimental quantum sim- ulation ...
  29. [29]
  30. [30]
    Quantum Field Theory - Stanford Encyclopedia of Philosophy
    Jun 22, 2006 · Historically, QFT resulted from the successful reconciliation of QM and SRT. In order to understand the initial problem one has to realize that ...What is QFT? · The Basic Structure of the... · Axiomatic Reformulations of...
  31. [31]
    [1812.10732] Introduction to the "second quantization" formalism for ...
    Dec 27, 2018 · This is a self-contained and hopefully readable account on the method of creation and annihilation operators (also known as the Fock space representation)
  32. [32]
    Jordan-Wigner Fermionization and the Theory of Low-Dimensional ...
    Dec 27, 2008 · The aim of this chapter is to review the application of the Jordan-Wigner fermionization technique for calculating dynamic quantities of low-dimensional ...