Fact-checked by Grok 2 weeks ago

Bogoliubov transformation

The Bogoliubov transformation is a linear canonical transformation in quantum mechanics and quantum field theory that mixes creation and annihilation operators to preserve the canonical commutation (for bosons) or anticommutation (for fermions) relations, enabling the diagonalization of quadratic Hamiltonians in interacting many-body systems. For bosonic modes, it typically takes the form \hat{b}_k = u_k \hat{a}_k + v_k \hat{a}^\dagger_{-k} with the condition |u_k|^2 - |v_k|^2 = 1, while for fermionic modes in contexts like superconductivity, it is \gamma_{k\uparrow} = u_k c_{k\uparrow} + v_k c^\dagger_{-k\downarrow} with |u_k|^2 + |v_k|^2 = 1. This transformation introduces quasiparticle operators that describe collective excitations, simplifying the treatment of phenomena where interactions lead to phenomena like pairing or condensation. Originally developed by Nikolai N. Bogoliubov in his 1947 microscopic theory of for interacting gases, the provided a method to approximate the and low-energy excitations of liquid helium-4 by accounting for quantum depletion of the . It was subsequently generalized and applied by Bogoliubov and collaborators, and independently by John George Valatin, in to the microscopic theory of (also known as the Bogoliubov–Valatin ), where it diagonalizes the Bardeen-Cooper-Schrieffer (BCS) to reveal the energy gap and paired electron states. Beyond these foundational roles, the Bogoliubov transformation has broad applications across quantum physics. In Bose-Einstein condensates, it describes Bogoliubov quasiparticles as the elementary excitations above the condensate ground state. In , it generates squeezed states and enables the analysis of parametric down-conversion processes. It also plays a crucial role in relativistic , such as in the and , where it relates inertial and accelerated observers' vacua through mode mixing in curved spacetimes.

General Mathematical Formulation

Canonical Transformations

The Bogoliubov transformation is defined as a linear isomorphism between sets of creation and annihilation operators that preserves the canonical (anti)commutation relations of bosonic or fermionic systems. For bosons, if \alpha and \alpha^\dagger satisfy [\alpha, \alpha^\dagger] = 1, the transformed operators \beta and \beta^\dagger take the general form \beta = u \alpha + v \alpha^\dagger and \beta^\dagger = u^* \alpha^\dagger + v^* \alpha, where u and v are complex coefficients. To ensure [\beta, \beta^\dagger] = 1, the condition |u|^2 - |v|^2 = 1 must hold. Similarly, for fermions, if c and c^\dagger satisfy \{c, c^\dagger\} = 1, the transformation \beta = u c + v c^\dagger preserves the anticommutation relation \{\beta, \beta^\dagger\} = 1 under the condition |u|^2 + |v|^2 = 1. These conditions are essential for maintaining the algebraic structure underlying quantum statistics, ensuring the transformation is canonical and induces a unitary evolution in the corresponding Fock space. For bosonic systems, the relation |u|^2 - |v|^2 = 1 reflects the symplectic nature of the transformation in phase space, where the operators correspond to position and momentum variables, preserving the Poisson bracket structure in the classical limit and the Heisenberg uncertainty principle quantum mechanically. In fermionic cases, the condition |u|^2 + |v|^2 = 1 aligns with the orthogonal group structure, guaranteeing unitarity and the Pauli exclusion principle. Both ensure the transformation can be implemented by a unitary operator on the Hilbert space, avoiding inconsistencies in expectation values or correlation functions. The Bogoliubov transformation was introduced by Nikolai Bogoliubov in his seminal work on the microscopic theory of , where it was used to describe collective excitations in weakly interacting gases. This framework has since become foundational in many-body quantum physics, enabling the treatment of paired states in both bosonic and fermionic systems.

Unified Matrix Description

The Bogoliubov transformation for multimode systems can be expressed in a compact notation by considering the column of operators \boldsymbol{\alpha} = \begin{pmatrix} \mathbf{a} \\ \mathbf{a}^\dagger \end{pmatrix}, where \mathbf{a} = (a_1, \dots, a_N)^T and \mathbf{a}^\dagger = (a_1^\dagger, \dots, a_N^\dagger)^T are the annihilation and creation operators for N modes, respectively. The transformed operators are given by \boldsymbol{\beta} = W \boldsymbol{\alpha}, where \boldsymbol{\beta} = \begin{pmatrix} \mathbf{b} \\ \mathbf{b}^\dagger \end{pmatrix} and W is a $2N \times 2N of the block form W = \begin{pmatrix} U & V \\ V^* & U^* \end{pmatrix}, with U and V being N \times N complex matrices. For the bosonic case, preservation of the canonical commutation relations [\beta_i, \beta_j^\dagger] = \delta_{ij} and [\beta_i, \beta_j] = 0 requires W to satisfy the paraunitary condition W \Sigma W^\dagger = \Sigma, where \Sigma = \operatorname{diag}(I_N, -I_N) is the indefinite . This condition implies the block relations U U^\dagger - V V^\dagger = I_N, U V^\dagger - V U^\dagger = 0, and their conjugates, ensuring the transformation belongs to the \mathrm{Sp}(2N, \mathbb{C}). Taking the determinant of the paraunitary condition yields \det(W) \det(W^\dagger) = 1, and given the structure of W in the real symplectic subgroup, the normalization follows as \det(W) = 1. In the fermionic case, the transformation preserves the canonical anticommutation relations \{\beta_i, \beta_j^\dagger\} = \delta_{ij} and \{\beta_i, \beta_j\} = 0 under the unitary condition W^\dagger W = I_{2N}, with the matrix form adjusted to W = \begin{pmatrix} U & V \\ -V^* & U^* \end{pmatrix} to ensure consistency (note the minus sign, conventional in paired systems like ). This leads to the block relations U U^\dagger + V V^\dagger = I_N and U V^\dagger - V U^\dagger = 0 (with conjugates), corresponding to the \mathrm{U}(2N) in the complex doubled space, preserving the structure. The normalization involves \det(W) = \det(U + i V) \det(U - i V) or equivalent, distinguishing proper transformations. The inversion of the transformation is given by \boldsymbol{\alpha} = W^{-1} \boldsymbol{\beta}. For the bosonic case, W^{-1} = \Sigma W^\dagger \Sigma, while for fermions, it follows from the unitary condition as W^{-1} = W^\dagger. This matrix formulation facilitates numerical implementations and algebraic manipulations, such as in the of quadratic Hamiltonians.

Bosonic Bogoliubov Transformations

Single-Mode Bosonic Case

In the single-mode bosonic case, the Bogoliubov transformation provides a canonical mapping between two sets of bosonic operators, mixing for a single mode. Specifically, the transformed annihilation operator is given by
b = u a + v a^\dagger,
where a and a^\dagger are the original bosonic operators satisfying the commutation relation [a, a^\dagger] = 1, and u, v \in \mathbb{C} are complex coefficients obeying the condition |u|^2 - |v|^2 = 1. This ensures the transformed operators b and b^\dagger also satisfy the bosonic commutation relation [b, b^\dagger] = 1. To verify this, compute
[b, b^\dagger] = u u^* [a, a^\dagger] + v v^* [a^\dagger, a] + u v^* [a, a] + v u^* [a^\dagger, a^\dagger] = |u|^2 - |v|^2 = 1,
since cross terms vanish under the original commutation rules.
The coefficients u and v can be parametrized using to satisfy the normalization condition explicitly:
u = \cosh \theta, \quad v = \sinh \theta \, e^{i \phi},
where \theta \geq 0 is a real squeezing and \phi is a angle. This form arises naturally in the context of quadratic bosonic Hamiltonians and highlights the transformation's connection to squeezing operations in . The inverse transformation is then a = u b - v b^\dagger, preserving the structure.
The state in the new basis, denoted |0'\rangle, is defined by the condition b |0'\rangle = 0. Expressed in the original Fock basis, this state takes the form of a squeezed :
|0'\rangle = \frac{1}{\sqrt{\cosh \theta}} \exp\left( \frac{1}{2} e^{i \phi} \tanh \theta \, (a^\dagger)^2 \right) |0\rangle,
where |0\rangle is the original . More explicitly in the Fock basis (for general phase \phi), it expands as
|0'\rangle = (\cosh \theta)^{-1/2} \sum_{n=0}^{\infty} \frac{ (\tanh \theta \, e^{i \phi})^n }{ \sqrt{2^{2n} (n!)^2 } } \sqrt{(2n)!} \, |2n\rangle.
This state represents a basic example of a squeezed , where quadrature fluctuations are reduced in one direction at the expense of the other, enabling applications in precision measurements.

Multimode Bosonic Case

In the multimode bosonic case, the Bogoliubov transformation generalizes the single-mode formulation to an arbitrary number of coupled bosonic modes, capturing collective excitations such as those in extended . This extension is essential for describing phenomena involving inter-mode interactions, where the transformation mixes and operators across multiple modes while preserving bosonic commutation relations [ \vec{b}_i, \vec{b}_j^\dagger ] = [\delta_{ij}](/page/Delta). The multimode transformation is expressed in vector form as \vec{b} = U \vec{a} + V \vec{a}^\dagger, where \vec{a} and \vec{b} are vectors of annihilation operators for the original and transformed modes, respectively, and U, V are complex matrices of appropriate dimensions. To maintain the canonical commutation relations, these matrices must satisfy the bosonic symplectic conditions: U U^\dagger - V V^\dagger = I and U V^T = V U^T. These orthogonality relations ensure the transformation is canonical and unitary in the extended phase space, allowing for the diagonalization of multimode quadratic forms. A representative example is the two-mode squeezing transformation in , where the matrices U and V mix operators from two spatial or temporal modes to generate entangled squeezed states, such as the two-mode squeezed vacuum |\psi\rangle = \sum_n \frac{(\tanh r)^n}{\cosh r} |n\rangle_a |n\rangle_b, with squeezing parameter r. This transformation, parameterized by U = \cosh r \, I and V = \sinh r \, \sigma (where \sigma couples the modes), produces non-classical correlations useful for quantum information tasks. In inhomogeneous bosonic systems, such as trapped gases, the multimode amplitudes are determined by the Bogoliubov-de Gennes (BdG) equations, which form a coupled set of eigenvalue problems for the wavefunctions u_j(\mathbf{r}) and v_j(\mathbf{r}): \begin{pmatrix} \mathcal{L} & \mathcal{M} \\ -\mathcal{M}^* & -\mathcal{L}^* \end{pmatrix} \begin{pmatrix} u_j \\ v_j \end{pmatrix} = E_j \begin{pmatrix} u_j \\ v_j \end{pmatrix}, where \mathcal{L} and \mathcal{M} incorporate the mean-field potential and interactions, yielding the matrices U and V from the eigenvectors. These equations extend the uniform case to spatially varying densities, enabling the study of localized excitations. Numerically, the matrices U and V are obtained by solving the generalized eigenvalue problem associated with the system's dynamical matrix, typically using iterative methods like the for large mode numbers to compute the positive-energy eigenvectors while enforcing the constraints. This approach scales efficiently for systems up to hundreds of modes, providing the transformation parameters for simulation of dynamics.

Fermionic Bogoliubov Transformations

Single-Mode Fermionic Case

The single-mode fermionic Bogoliubov transformation mixes the and operators of a single fermionic mode to diagonalize quadratic Hamiltonians with particle-hole . The is expressed as \gamma = u \, c + v \, c^\dagger, where c and c^\dagger are the original fermionic operators satisfying \{c, c^\dagger\} = 1 and \{c, c\} = 0, and u, v are coefficients obeying the |u|^2 + |v|^2 = 1. This condition arises from requiring the transformed operator to preserve the canonical anticommutation relation \{\gamma, \gamma^\dagger\} = 1, ensuring \gamma behaves as a valid fermionic quasiparticle operator. The transformation exhibits a close relation to Majorana fermions through real linear combinations of the operators. With u and v taken as real, the combination \gamma + \gamma^\dagger = (u + v)(c + c^\dagger) is proportional to a real Majorana operator, defined as the Hermitian part \eta_1 = (c + c^\dagger)/\sqrt{2}, while the imaginary part involves \eta_2 = -i(c - c^\dagger)/\sqrt{2}. These Majorana operators satisfy \{\eta_i, \eta_j\} = \delta_{ij} and \eta_i^\dagger = \eta_i, underscoring the inherent particle-hole mixing in the single-mode case. The operators obey the fermionic anticommutation relations \{\gamma, \gamma^\dagger\} = 1 and \{\gamma, \gamma\} = 0, which follow directly from the and the original operators' statistics. The corresponding state, annihilated by \gamma (i.e., \gamma |\Omega\rangle = 0), can be interpreted in the field-theoretic sense as a filled , where all negative-energy states below the are occupied to ensure stability. In the context of pairing mechanisms, this manifests as a coherent paired state, a superposition that correlates the unoccupied and occupied configurations of the mode. In many theoretical models, the phases of u and v are chosen to be real (e.g., u = \cos\theta, v = \sin\theta for some mixing angle \theta) to exploit symmetries in the Hamiltonian and simplify the diagonalization, aligning the transformation with time-reversal or particle-hole invariance. This single-mode formulation serves as the foundational building block and can be represented in a unified 2×2 matrix form for extension to multimode scenarios.

Multimode Fermionic Case

In the multimode fermionic case, the Bogoliubov transformation generalizes to a linear mixing of across multiple fermionic modes, preserving the canonical anticommutation relations \{ \gamma_i, \gamma_j^\dagger \} = \delta_{ij} and \{ \gamma_i, \gamma_j \} = 0. The transformation takes the vector form \vec{\gamma} = U \vec{c} + V \vec{c}^\dagger, where \vec{c} and \vec{c}^\dagger are vectors of original annihilation and creation operators for N modes, and U, V are N \times N matrices satisfying the orthonormality conditions U U^\dagger + V V^\dagger = I and U V^T + V U^T = 0 to ensure the transformed operators \vec{\gamma} obey fermionic . This multimode framework is central to the Bogoliubov-de Gennes (BdG) formalism, which describes quasiparticle excitations in inhomogeneous superconductors by coupling electron-like and hole-like components across momentum-space modes. In the BdG approach, the mean-field Hamiltonian for a superconductor is expressed in a doubled Nambu-Gorkov space, leading to an eigenvalue problem that diagonalizes the quadratic form via the multimode transformation. The resulting BdG equations are \begin{pmatrix} H_0 & \Delta \\ \Delta^\dagger & -H_0^* \end{pmatrix} \begin{pmatrix} u_k \\ v_k \end{pmatrix} = E_k \begin{pmatrix} u_k \\ v_k \end{pmatrix}, where H_0 is the single-particle Hamiltonian (e.g., kinetic energy plus potential), \Delta is the pairing potential matrix coupling modes k and k', and u_k, v_k are the electron and hole wavefunction components for quasiparticle mode k. The eigenvalue spectrum features positive energies E_k > 0 for particle-like excitations and negative energies -E_k < 0 for hole-like excitations, with the physical quasiparticle spectrum given by the positive branch due to particle-hole symmetry in the fermionic system. This structure arises because the transformation mixes particle and hole sectors, yielding a doubled set of eigenvalues symmetric around zero, and the ground state is the vacuum of the \gamma operators annihilating all positive-energy modes. A representative example is singlet pairing in s-wave superconductors, where the off-diagonal pairing matrix V couples opposite momenta such that V_{k, -k} \neq 0 while V_{k, k'} = 0 for k' \neq -k, reflecting time-reversal symmetry and momentum conservation in the uniform case. This form simplifies the BdG equations to decoupled pairs of k and -k modes, yielding quasiparticle energies E_k = \sqrt{\xi_k^2 + |\Delta_k|^2}, where \xi_k is the normal-state dispersion and \Delta_k the s-wave gap. The single-mode fermionic case can be viewed as a local approximation to this multimode structure in position space.

Diagonalization of Quadratic Hamiltonians

Bosonic Hamiltonians

Quadratic bosonic Hamiltonians arise in many-body systems and quantum optics, describing non-interacting quasiparticles after applying a . In the multimode case, such a Hamiltonian takes the general form H = \vec{a}^\dagger H_0 \vec{a} + \frac{1}{2} \left( \vec{a}^\dagger M \vec{a}^\dagger + \vec{a} M^* \vec{a} \right), where \vec{a} is the column vector of annihilation operators for N modes satisfying [\vec{a}_i, \vec{a}_j^\dagger] = \delta_{ij}, H_0 is an N \times N Hermitian matrix, and M is an N \times N symmetric complex matrix. This form captures both kinetic-like terms (diagonal in creation-annihilation) and pairing terms (anomalous contributions mixing creation with creation or annihilation with annihilation). To diagonalize H using a bosonic Bogoliubov transformation, one introduces new operators \vec{b} via \vec{b} = U \vec{a} + V \vec{a}^\dagger and \vec{b}^\dagger = \vec{a}^\dagger U^\dagger + \vec{a} V^\dagger, where U and V are N \times N matrices satisfying the bosonic canonical commutation relations U U^\dagger - V V^\dagger = I and U V^T = V U^T (symplectic conditions). Substituting into H yields a diagonal form H = \sum_k E_k b_k^\dagger b_k + C, where E_k > 0 are the energies and C is a constant ( shift), provided the original is stable (i.e., bounded from below). The matrices U and V are determined by solving the eigenvalue problem for the doubled $2N \times 2N matrix \mathcal{H} = \begin{pmatrix} H_0 & M \\ M^* & H_0 \end{pmatrix}, assuming H_0 is real symmetric for simplicity in many cases. The procedure involves finding a paraunitary matrix W (satisfying W^\dagger \Sigma W = \Sigma with \Sigma = \operatorname{diag}(I, -I)) such that W^\dagger \mathcal{H} W = \operatorname{diag}(E, -E), where E = \operatorname{diag}(E_1, \dots, E_N) contains the positive eigenvalues. The columns of W = \begin{pmatrix} U \\ V \end{pmatrix} corresponding to positive eigenvalues define the transformation coefficients, ensuring the operators \vec{b} obey the same commutation relations as \vec{a}. The eigenvalues E_k must be real and positive for the spectrum to be bounded below, with the constant given by C = -\frac{1}{2} \sum_k E_k. This leverages Williamson's for of positive-definite forms. A representative example is the single-mode with a squeezing term: H = \omega a^\dagger a + \frac{\kappa}{2} (a^{\dagger 2} + a^2), where \omega > 0 and |\kappa| < \omega for stability. Here, H_0 = \omega and M = \kappa (real symmetric). The doubled matrix is \mathcal{H} = \begin{pmatrix} \omega & \kappa \\ \kappa & \omega \end{pmatrix}, with eigenvalues \pm E where E = \sqrt{\omega^2 - \kappa^2}. The Bogoliubov transformation is b = u a + v a^\dagger with u = \cosh r, v = \sinh r, and \tanh 2r = \kappa / \omega, yielding H = E b^\dagger b - \frac{E}{2}. This illustrates how the transformation mixes to uncouple the modes into free quasiparticles.

Fermionic Hamiltonians

Quadratic fermionic Hamiltonians arise in models of and fermionic , where interactions lead to pairing terms that couple particle . A example is the BCS-like Hamiltonian, expressed as
H = \sum_k \varepsilon_k c_k^\dagger c_k + \sum_k \left( \Delta_k c_k^\dagger c_{-k}^\dagger + \Delta_k^* c_{-k} c_k \right),
where c_k^\dagger and c_k are fermionic satisfying anticommutation relations \{c_k, c_l^\dagger\} = \delta_{kl}, \varepsilon_k represents the single-particle relative to the , and \Delta_k is the pairing potential. This form captures the essential physics of Cooper pair formation while remaining quadratic in the operators, allowing exact diagonalization.
To diagonalize this , a fermionic Bogoliubov transformation is applied, introducing operators \gamma_k = u_k c_k + v_k c_{-k}^\dagger and \gamma_{-k} = u_k c_{-k} - v_k c_k^\dagger, with real coefficients satisfying u_k^2 + v_k^2 = 1 to preserve anticommutation relations. Substituting these into the Hamiltonian yields the Bogoliubov-de Gennes (BdG) equations as an eigenvalue problem for the coefficients:
\begin{pmatrix} \varepsilon_k & \Delta_k \\ \Delta_k^* & -\varepsilon_k \end{pmatrix} \begin{pmatrix} u_k \\ v_k \end{pmatrix} = E_k \begin{pmatrix} u_k \\ v_k \end{pmatrix}.
The solutions give eigenvalues E_k = \pm \sqrt{\varepsilon_k^2 + |\Delta_k|^2}, with corresponding eigenvectors determining u_k and v_k. These \gamma_k diagonalize the Hamiltonian, transforming it into
H = \sum_k E_k \gamma_k^\dagger \gamma_k + \text{constant},
where the constant term arises from the ground-state energy shift, typically -\sum_k E_k / 2.
The BdG spectrum features both positive and negative eigenvalues due to the structure of the pairing terms, reflecting the electron-hole mixing in the excitations. However, particle-hole redundancy in the fermionic description implies that negative-energy states are not independent but correspond to the positive-energy excitations of the conjugate modes; specifically, if (u_k, v_k, E_k) is an eigenstate, so is (v_k^*, -u_k^*, -E_k) for the particle-hole . To obtain a complete set of positive-energy operators, only the positive branch E_k \geq 0 is retained, ensuring the diagonal form expresses the solely in terms of non-negative excitation energies above the . This redundancy is a hallmark of the BdG formalism, preventing double-counting of in the doubled Nambu space. In multimode scenarios, such as inhomogeneous superconductors, the transformation generalizes to position-dependent coefficients, but the core algebraic structure remains analogous to the momentum-space case.

Applications and Physical Interpretations

Bose-Einstein Condensates

The mean-field description of a Bose-Einstein condensate (BEC) is provided by the Gross-Pitaevskii equation, which models the condensate \psi(\mathbf{r}, t) for a dilute, weakly interacting bosonic gas as i \hbar \frac{\partial \psi}{\partial t} = \left[ -\frac{\hbar^2}{2m} \nabla^2 + V(\mathbf{r}) + g |\psi|^2 \right] \psi, where m is the atomic mass, V(\mathbf{r}) is the external trapping potential, and g = 4\pi \hbar^2 a / m is the interaction parameter with s-wave scattering length a. For a uniform condensate at zero temperature, the ground state is \psi = \sqrt{n} e^{-i \mu t / \hbar} with chemical potential \mu = g n, where n is the condensate density. To account for low-energy excitations around this , the Bogoliubov approximation treats fluctuations \delta \psi(\mathbf{r}, t) about the , expanding the field operator as \hat{\Psi} = \psi + \delta \hat{\Psi}. In the linear regime for a homogeneous system, these fluctuations take the form \delta \psi(\mathbf{r}, t) = u_k e^{i \mathbf{k} \cdot \mathbf{r} - i E_k t / \hbar} + v_k^* e^{-i \mathbf{k} \cdot \mathbf{r} + i E_k t / \hbar}, where u_k and v_k are Bogoliubov amplitudes satisfying the |u_k|^2 - |v_k|^2 = 1. This , combined with a quadratic expansion of the , leads to the diagonalization via bosonic Bogoliubov transformations, yielding excitations. The resulting excitation spectrum is E_k = \sqrt{\epsilon_k (\epsilon_k + 2 g n)}, where \epsilon_k = \hbar^2 k^2 / 2m is the free-particle energy. At low momenta (k \to 0), this disperses linearly as E_k \approx \hbar c k with sound velocity c = \sqrt{g n / m}, reflecting the phonon regime characteristic of . At higher momenta, it approaches the free-particle form \epsilon_k, marking the transition to particle-like excitations. This spectrum emerges from solving the coupled Bogoliubov-de Gennes equations for u_k and v_k. A key prediction is the quantum depletion of the , representing the fraction of particles excited out of the zero-momentum state due to interactions. The number of depleted particles is N_\mathrm{dep} = \int \frac{d^3 k}{(2\pi)^3} |v_k|^2, with |v_k|^2 = \frac{1}{2} \left( \frac{\epsilon_k + g n}{E_k} - 1 \right). Evaluating this integral yields N_\mathrm{dep}/N \approx \frac{8}{3\sqrt{\pi}} (n a^3)^{1/2}, which remains small (\ll 1) for dilute gases where the gas parameter n a^3 \ll 1, justifying the . Although originally formulated for superfluid helium, the Bogoliubov theory found direct application to dilute BECs following their experimental realization in trapped vapors of rubidium-87 and sodium-23 in 1995. These systems, with densities n \sim 10^{12} - 10^{15} cm^{-3} and temperatures near 100 nK, provided ideal conditions for testing the theory, with validations through measurements of the excitation spectrum via time-of-flight expansion and Bragg spectroscopy, as well as direct imaging of depletion effects.

Superconductivity and Fermionic Condensates

In the context of superconductivity, the Bogoliubov transformation is applied to fermionic systems to diagonalize the mean-field Hamiltonian derived from Bardeen-Cooper-Schrieffer (BCS) theory, which describes the pairing of electrons into Cooper pairs below a critical temperature. This transformation mixes creation and annihilation operators for electrons and holes, leading to a diagonal form of the Hamiltonian in terms of quasiparticle operators. The resulting quasiparticle dispersion relation is given by E_k = \sqrt{\epsilon_k^2 + |[\Delta](/page/Delta)|^2}, where \epsilon_k is the single-particle relative to the , and \Delta is the superconducting order parameter representing the pairing amplitude. The quasiparticles defined by the Bogoliubov provide a physical interpretation of excitations in the superconducting state: the \gamma_k^\dagger creates a broken , consisting of a quasiparticle above the and a quasihole below it, with an energy cost E_k that exhibits an |\Delta| at the . The is parameterized by coherence factors u_k and v_k, satisfying the normalization condition u_k^2 + v_k^2 = 1, which determine the amplitude of the particle-like and hole-like components, respectively; specifically, u_k^2 = \frac{1}{2} \left(1 + \frac{\epsilon_k}{E_k}\right) and v_k^2 = \frac{1}{2} \left(1 - \frac{\epsilon_k}{E_k}\right). These factors ensure that the is a coherent superposition of paired states, and the excitation spectrum reflects the stability of the paired phase. Self-consistency in BCS theory is enforced through the gap equation, which relates the order parameter to the thermal occupation of quasiparticles: \Delta = -V \sum_k u_k v_k \tanh\left(\frac{\beta E_k}{2}\right), with u_k v_k = \frac{|\Delta|}{2 E_k}, where V is the attractive pairing interaction strength (assumed constant near the Fermi surface), \beta = 1/(k_B T), and the sum is over momentum states; at zero temperature, \tanh(\beta E_k / 2) \to 1, simplifying the equation and yielding a finite \Delta only below the critical temperature. This equation is solved iteratively to determine the temperature dependence of the gap, confirming the exponential suppression of \Delta near T_c. The formalism of Bogoliubov transformations in extends naturally to fermionic superfluids in ultracold atomic gases, where tunable interactions via Feshbach resonances allow exploration of the BCS-BEC crossover. Following the first experimental evidence of in resonantly interacting Fermi gases of ^6Li atoms in 2003–2004, these systems have been used to realize and probe paired fermionic condensates, with the same quasiparticle and describing the superfluid across interaction strengths from weak (BCS-like) to strong .

Other Contexts

In , Bogoliubov transformations underpin the generation of , which reduce in one at the expense of the other, enabling applications in precision measurements and processing. For single-mode squeezing, the acts as a Bogoliubov transformation on the bosonic , transforming the into a squeezed vacuum state with variance (\Delta X)^2 = e^{-2r} in the squeezed quadrature, where r is the squeezing . This formulation links directly to the single-mode bosonic case, where the transformation mixes creation and annihilation operators to diagonalize the of parametric down-conversion processes in nonlinear media. Experimental realization of squeezed light was first achieved using optical parametric oscillators, demonstrating below the . Two-mode squeezing extends this to entangled states across two optical modes, described by a multimode Bogoliubov transformation that correlates the quadratures, as in the H = i \hbar \kappa (a^\dagger b^\dagger - a b), where a, b are mode s and \kappa is the coupling strength; this produces Einstein-Podolsky-Rosen-like entanglement useful for and . In , Bogoliubov transformations describe particle creation from the in an expanding , particularly during the , where time-dependent metrics lead to mixing of positive and negative frequency modes. The transformation relates "in" and "out" vacua via coefficients \alpha_k and \beta_k, with the particle number density given by |\beta_k|^2, quantifying the production of quanta that seed fluctuations. This mechanism, first derived for a conformally flat , applies to massless fields in Friedmann-Lemaître-Robertson-Walker metrics and serves as an analog to in curved . During , the rapid expansion amplifies quantum fluctuations, with \beta_k determined by solving the mode equation \ddot{\phi}_k + 3H \dot{\phi}_k + (k^2/a^2 + m^2) \phi_k = 0, where H is the Hubble parameter and a(t) the scale factor, leading to a thermal-like of created particles. Recent developments have extended Bogoliubov transformations to (QED) for parametric amplifiers, where post-2010 experiments utilize Josephson junctions to modulate frequencies, implementing bosonic transformations that amplify signals while adding minimal noise. In these systems, a driven superconducting undergoes a time-dependent Bogoliubov transformation, converting fluctuations into squeezed photons with gains exceeding 20 , as demonstrated in arrays of superconducting quantum interference devices. This enables quantum-limited amplification for readout in architectures. In topological superconductors, fermionic Bogoliubov transformations within the Bogoliubov-de Gennes formalism reveal Majorana zero modes at defects or edges, which are self-conjugate quasiparticles robust against decoherence. These modes emerge in one-dimensional p-wave superconductors or proximitized nanowires, where the H = \sum_k (c_k^\dagger, c_{-k}) \begin{pmatrix} \xi_k & \Delta_k \\ \Delta_k^* & -\xi_k \end{pmatrix} \begin{pmatrix} c_k \\ c_{-k}^\dagger \end{pmatrix} yields zero-energy solutions at topological phase transitions, protected by particle-hole symmetry and proposed for fault-tolerant quantum bits. Analog models of the Hawking effect in sonic black holes formed within Bose-Einstein condensates (BECs) employ Bogoliubov theory to predict and observe thermal phonon emission at the sonic horizon, where the flow velocity exceeds the . In these quasi-one-dimensional BECs, the low-energy excitations follow a relativistic near the horizon, and the Bogoliubov transformation between subsonic and supersonic regions yields a with T = \hbar c_s / (2\pi k_B \kappa), where c_s is the and \kappa the surface gravity analog. Experiments in elongated BEC traps have detected correlated pairs across the horizon, confirming Hawking-like with effective temperatures around 100 nK, providing a laboratory testbed for quantum field effects in curved .

References

  1. [1]
  2. [2]
    Bogolyubov's theory of superfluidity | Physics of Particles and Nuclei
    Nov 26, 2010 · References. N. N. Bogolyubov, J. Phys. USSR 11, 23 (1947). MathSciNet Google Scholar. N. N. Bogolyubov, Bull. Mosc. State Univ. 7, 43 (1947).
  3. [3]
  4. [4]
    1 Introduction
    Sep 1, 2025 · The Bogoliubov transformation is a powerful tool to find normal modes and diagonalize interacting hamiltonians in many-body physics [1] . It ...
  5. [5]
    Virasoro-Bogoliubov Transformations in Two Modes - arXiv
    The Bogoliubov transformation is a fundamentally important physical tool in quantum field theory. This transformation essentially exhibits non-perturbative ...
  6. [6]
    On the theory of superfluidity - Inspire HEP
    On the theory of superfluidity. N.N. Bogolyubov(. Steklov Math. Inst., Moscow. ) 1947. 14 pages ... 11 (1947) 77-90. cite claim. reference search 544 citations ...
  7. [7]
    None
    Summary of each segment:
  8. [8]
    [PDF] M. Stone Jan 2017 and Dec 2019 Bose Bogoliubov transformations ...
    Let us write the Bogoluibov transformation as aa∗ = U V ∗. V U∗ bb∗ , where U, V are n-by-n complex matrices. The commutation relations are preserved if ...
  9. [9]
    [PDF] Bogoliubov transformations
    Consider a Fock bosonic or fermionic representation with an ar- bitrary number of degrees of freedom. We say that a self-adjoint operator H is a quadratic ...<|control11|><|separator|>
  10. [10]
    [PDF] arXiv:2306.00795v1 [quant-ph] 1 Jun 2023
    Jun 1, 2023 · We call all maps that preserve the commutation relations of a fermionic-anyon algebra as canonical transformations, as in the following ...<|control11|><|separator|>
  11. [11]
    [PDF] Bogoliubov Transformation, Group Structure
    Bogoliubov Transformation, Group Structure. Any F-even unitary matrix in 2M will correspond to some Bogoliubov transformation. The set of such matrices forms ...
  12. [12]
  13. [13]
  14. [14]
    [quant-ph/0109020] General Multimode Squeezed States - arXiv
    Sep 4, 2001 · By the complex multimode Bogoliubov transformation, we obtain the general forms of squeeze operators and squeezed states including squeezed ...
  15. [15]
    Multiparticle quantum interference in Bogoliubov bosonic ...
    Oct 22, 2021 · We explore the multiparticle transition probabilities in Gaussian unitaries effected by a two-mode Bogoliubov bosonic transformation on the modeAbstract · Article Text · INTRODUCTION · MODEL AND DERIVATIONS
  16. [16]
  17. [17]
  18. [18]
  19. [19]
    [PDF] ON THE BOGOLIUBOV-DE GENNES EQUATIONS
    MacArthur Foundation. Page 2. TQT - 1. On the Bogoliubov-de Gennes equations. ... superconductors, has been based on the Bogoliubov-de. Gennes (generalized ...
  20. [20]
    [PDF] Bosonic quadratic Hamiltonians - arXiv
    Jan 1, 2017 · (In the bosonic context the term “Bogoliubov transformations” is usually meant to denote “symplectic transformations”). The abstract approach ...
  21. [21]
    [PDF] Superconductors and the Bogoliubov–de Gennes “trick”
    Mar 10, 2016 · called the Bogoliubov–de Gennes formalism, that involves a redundant repre- sentation of the states. We begin by rewriting the Hamiltonian ...Missing: multimode | Show results with:multimode
  22. [22]
    Bogoliubov theory for dilute Bose gases: The Gross-Pitaevskii regime
    Aug 2, 2019 · In 1947, Bogoliubov suggested a heuristic theory to compute the excitation spectrum of weakly interacting Bose gases.
  23. [23]
    [PDF] Bogoliubov theory for dilute bose gases: the Gross-Pitaevskii Regime
    Mar 19, 2019 · Abstract. In 1947 Bogoliubov suggested a heuristic theory to compute the excitation spec- trum of weakly interacting Bose gases.
  24. [24]
    The Bogoliubov model of weakly imperfect Bose gas - ScienceDirect
    We present a systematic account of known rigorous results about the Bogoliubov model of weakly imperfect Bose gas (WIBG).
  25. [25]
    Quantum Depletion of a Homogeneous Bose-Einstein Condensate
    Nov 7, 2017 · In 1947, Bogoliubov developed a theory that explains the microscopic origin of such interaction-driven, quantum depletion of a BEC [7]
  26. [26]
    Observation of Bose-Einstein Condensation in a Dilute Atomic Vapor
    The condensate fraction first appeared near a temperature of 170 nanokelvin and a number density of 2.5 × 1012 per cubic centimeter and could be preserved for ...
  27. [27]
    June 5, 1995: First Bose Einstein Condensate
    Jun 1, 2004 · The world's first BEC was achieved at 10:54 AM on June 5, 1995 in a laboratory at JILA, a joint institute of University of Colorado, Boulder, and NIST.