Fact-checked by Grok 2 weeks ago

Fock state

A Fock state is a in the of a many-particle system, characterized by a definite, fixed number of (such as bosons or fermions) occupying specific single-particle modes, typically denoted as |n\rangle where n is the occupation number. These states serve as the for the , which encompasses all possible particle numbers from zero (the vacuum state |0\rangle) to arbitrarily large values, enabling the description of systems with variable particle counts in . The concept was introduced by Soviet physicist in 1932 as part of his foundational work on , providing a formalism to handle identical particles without explicit labeling, thus avoiding issues in for . In the context of bosonic systems, such as photons in , Fock states are eigenstates of the number \hat{n} = \hat{a}^\dagger \hat{a}, with eigenvalue n, and are generated by applying the creation \hat{a}^\dagger successively to the : |n\rangle = \frac{(\hat{a}^\dagger)^n}{\sqrt{n!}} |0\rangle. They exhibit highly nonclassical features, including sub-Poissonian photon statistics and perfect number squeezing, distinguishing them from coherent states that approximate classical light. For fermions, Fock states respect the Pauli exclusion principle, limiting n to 0 or 1 per mode, which is crucial for modeling electronic structures in condensed matter physics. Fock states are fundamental in and technologies, enabling precise control of particle numbers for applications like quantum metrology, where their well-defined occupancy enhances measurement sensitivity beyond classical limits, and in quantum simulation of many-body Hamiltonians using photonic or superconducting circuits. Experimental generation of Fock states, particularly for photons, has been achieved through techniques such as parametric down-conversion or , with demonstrations up to occupation numbers like n = 15 in microwave cavities. More recent advances have generated states approaching 100 photons as of 2024. Their role extends to fundamental tests of , including entanglement generation and violation of Bell inequalities with number states.

Definition and Fundamentals

Definition

In second quantization, Fock states represent the fundamental basis for describing many-body quantum systems of identical particles, serving as simultaneous eigenstates of the particle number operator associated with each single-particle mode. These states capture the occupation numbers of particles in specific modes, providing a natural framework for handling indistinguishable particles without explicit symmetrization or antisymmetrization of wave functions. Mathematically, a Fock state for a single mode is denoted as |n\rangle, where it satisfies the eigenvalue equation \hat{a}^\dagger \hat{a} \, |n\rangle = n \, |n\rangle, with \hat{a}^\dagger and \hat{a} being the creation and annihilation operators, respectively, and n the occupation number (an integer ≥ 0 for bosons or 0 or 1 for fermions). For multiple modes, the state generalizes to |\{n_i\}\rangle, a product over modes labeled by occupation numbers n_i, forming a basis in the Fock space. This representation arises from the algebraic structure of second quantization, where operators act to change occupation numbers while preserving commutation relations appropriate to the particle statistics. The specific form of the operators and state normalization differs for bosons (commutation relations) and fermions (anticommutation relations), as detailed in subsequent sections. The concept is named after Soviet physicist , who introduced it in the early 1930s as part of developing the formalism of for and many-particle systems. Fock's work established the as the arena for these states, enabling a unified treatment of quantum fields where particle creation and annihilation are intrinsic. A key distinction exists between Fock states in representations with fixed particle number—where the is confined to sectors of definite total particles N—and those allowing variable particle number, as in the full , which is a over all N and permits superpositions across different particle counts. This flexibility is essential for systems like quantum fields or grand canonical ensembles, where particle number fluctuates.

Vacuum State

The vacuum state, denoted as |0\rangle, is defined as the unique state in the that is annihilated by the annihilation operator: \hat{a} |0\rangle = 0. This property ensures that no particles can be removed from the , establishing it as the foundational element of the Fock basis. In the of the system, this state is normalized such that \langle 0 | 0 \rangle = 1, and its uniqueness follows from being the only joint eigenvector of the relevant operators, including the and , corresponding to the configuration. Physically, the vacuum state represents the lowest energy configuration of the quantum system, with zero particles occupying the mode, as it is the eigenstate of the particle number operator \hat{n} = \hat{a}^\dagger \hat{a} with eigenvalue zero. Despite containing no real particles, the vacuum is not devoid of activity; it exhibits quantum fluctuations arising from the Heisenberg uncertainty principle, manifesting as temporary virtual particle-antiparticle pairs that contribute to zero-point energy. This serves as the starting point for constructing all higher Fock states through successive applications of the creation operator \hat{a}^\dagger. In bosonic systems, the n-particle state is generated as |n\rangle = \frac{(\hat{a}^\dagger)^n}{\sqrt{n!}} |0\rangle. In quantum field theories and , this role underscores the vacuum's centrality as the reference from which excitations build the full spectrum of particle occupations.

Bosonic Fock States

Boson Operators

In the , the annihilation operator \hat{a} and the creation operator \hat{a}^\dagger act on single-mode states and satisfy the [\hat{a}, \hat{a}^\dagger] = 1. These operators, originally introduced by Paul Dirac in the context of the quantum harmonic oscillator and radiation theory, enable the description of systems where multiple particles can occupy the same state without restriction. The annihilation operator \hat{a} is non-Hermitian, meaning \hat{a} \neq \hat{a}^\dagger, with the creation operator serving as its such that (\hat{a}^\dagger)^\dagger = \hat{a}. This non-Hermiticity arises directly from the commutation relation, as a would satisfy [\hat{a}, \hat{a}] = 0, contradicting the bosonic algebra. The commutation relations extend to the creation operators as [\hat{a}^\dagger, \hat{a}^\dagger] = 0 and [\hat{a}, \hat{a}] = 0, preserving the bosonic symmetry. A key identity in the algebra is the definition of the number operator \hat{n} = \hat{a}^\dagger \hat{a}, which counts the occupation number of bosons in the mode and satisfies [\hat{a}, \hat{n}] = -\hat{a} and [\hat{a}^\dagger, \hat{n}] = \hat{a}^\dagger. This operator plays a central role in the Hamiltonian for free bosonic fields, H = \hbar \omega (\hat{n} + 1/2), where \omega is the mode frequency. The normalized bosonic Fock states are generated by successive application of the creation operator on the vacuum state |0\rangle, defined such that \hat{a} |0\rangle = 0 and \langle 0 | 0 \rangle = 1. Specifically, the n-particle state is |n\rangle = \frac{(\hat{a}^\dagger)^n}{\sqrt{n!}} |0\rangle, ensuring \langle m | n \rangle = \delta_{mn} through the commutation . This normalization follows from the recursive action \hat{a}^\dagger |n\rangle = \sqrt{n+1} |n+1\rangle.

Bosonic Basis States

In the single-mode case, bosonic Fock states, denoted as |n⟩ where n is the occupation number (n = 0, 1, 2, ...), are constructed by successively applying the bosonic creation operator \hat{a}^\dagger to the vacuum state |0⟩. The explicit form is given by |n\rangle = \frac{(\hat{a}^\dagger)^n}{\sqrt{n!}} |0\rangle, which ensures proper normalization, \langle n | n \rangle = 1, as the factor \sqrt{n!} accounts for the degeneracy arising from the indistinguishability of bosons. This construction relies on the bosonic commutation relations [\hat{a}, \hat{a}^\dagger] = 1, which allow unlimited occupation numbers unlike fermionic cases. These states form an for the of the bosonic mode, satisfying the orthogonality relation \langle m | n \rangle = \delta_{mn} for m ≠ n. Moreover, they constitute a complete set, as expressed by the resolution of the identity \sum_{n=0}^\infty |n\rangle \langle n| = \hat{1}, enabling the expansion of any state in the space. For multi-particle configurations, the bosonic Fock states exhibit symmetric wavefunctions under particle exchange, reflecting the indistinguishability of bosons; the many-body wavefunction is the permanent of the single-particle orbitals occupied according to the Fock vector. As an illustrative example, the two-particle state is |2⟩ = \frac{(\hat{a}^\dagger)^2}{\sqrt{2!}} |0\rangle = \frac{1}{\sqrt{2}} (\hat{a}^\dagger)^2 |0\rangle, where the normalization factor \frac{1}{\sqrt{2}} arises from the two indistinguishable ways to place two bosons in the mode.

Operator Actions on Bosonic States

In bosonic systems, the annihilation operator \hat{a} acts on a Fock state |n\rangle by reducing the particle number by one, yielding \hat{a} |n\rangle = \sqrt{n} |n-1\rangle. This action preserves the normalization of the states, as the coefficient \sqrt{n} accounts for the overlap between the adjacent number states. Similarly, the creation operator \hat{a}^\dagger increases the particle number by one, with \hat{a}^\dagger |n\rangle = \sqrt{n+1} |n+1\rangle. These operators function as ladder operators, systematically stepping up or down the energy eigenstates in the Fock basis while maintaining orthogonality and unit norm. The number \hat{n} = \hat{a}^\dagger \hat{a} measures the occupation number, acting diagonally on Fock states as \hat{n} |n\rangle = n |n\rangle. This eigenvalue equation confirms that each |n\rangle is an eigenstate of the particle count, with eigenvalue n, which is central to the definition of Fock states in bosonic . To illustrate, consider the state |0\rangle: the annihilation gives \hat{a} |0\rangle = 0, annihilating no particles and producing the zero vector, while \hat{a}^\dagger |0\rangle = |1\rangle generates the first excited state, and \hat{n} |0\rangle = 0 \cdot |0\rangle. For the single-particle state |1\rangle, \hat{a} |1\rangle = \sqrt{1} |0\rangle = |0\rangle, \hat{a}^\dagger |1\rangle = \sqrt{2} |2\rangle, and \hat{n} |1\rangle = 1 \cdot |1\rangle. These examples highlight the ladder-like progression, where repeated applications of \hat{a}^\dagger build higher Fock states from the , with coefficients ensuring proper normalization.

Fermionic Fock States

Fermion Operators

In fermionic , the creation \hat{a}^\dagger and annihilation \hat{a} for a single fermionic mode are defined to satisfy the anticommutation relations \{\hat{a}, \hat{a}^\dagger\} = \hat{a} \hat{a}^\dagger + \hat{a}^\dagger \hat{a} = 1 and \{\hat{a}, \hat{a}\} = \hat{a} \hat{a} + \hat{a} \hat{a} = 0, along with \{\hat{a}^\dagger, \hat{a}^\dagger\} = 0. These relations ensure that the operators generate states obeying , fundamentally differing from the bosonic case through the use of anticommutators rather than commutators. Like their bosonic counterparts, the fermionic operators are non-Hermitian, with \hat{a}^\dagger serving as the of \hat{a}, but the anticommutation imposes stricter constraints on state occupations. A key identity derived from the fundamental relations is \hat{a}^\dagger \hat{a} + \hat{a} \hat{a}^\dagger = 1, which highlights the binary nature of fermionic occupancy. The number operator \hat{n} = \hat{a}^\dagger \hat{a} thus has eigenvalues restricted to 0 or 1, as applying \hat{a}^\dagger twice to the vacuum yields zero due to the anticommutator \{\hat{a}^\dagger, \hat{a}^\dagger\} = 0. This algebraic structure enforces the at the operator level, limiting each single-particle mode to at most one , which is essential for constructing antisymmetric many-body wave functions in . Consequently, fermionic for a single mode span a two-dimensional , precluding higher occupation numbers inherent in bosonic systems.

Fermionic Basis States

In the single-mode case for fermions, the Fock space is two-dimensional, spanned by the vacuum state |0\rangle and the one-particle state |1\rangle = \hat{a}^\dagger |0\rangle, where \hat{a}^\dagger is the creation operator for that mode. Higher occupation numbers are impossible because applying the creation operator twice yields zero due to the fermionic anticommutation relations. This enforces the , limiting each mode to at most one particle. For multi-mode fermionic systems, the basis states are labeled by occupation numbers \{n_k\}, where each n_k = 0 or $1 for mode k, and the state is constructed as |\{n_k\}\rangle = \prod_k (\hat{a}^\dagger_k)^{n_k} |0\rangle, with normalization factor $1 / \sqrt{\prod_k n_k !}, which simplifies to 1 since n_k ! = 1 for n_k = 0, 1. These states form an for the fermionic , which is the over particle sectors with antisymmetrized tensor products. In the position representation, the multi-particle wavefunctions corresponding to these states are Slater determinants, ensuring full antisymmetry under particle exchange. For instance, a two-particle state in distinct modes is given by the determinant \frac{1}{\sqrt{2}} \det \begin{pmatrix} \phi_{k_1}(\mathbf{x}_1) & \phi_{k_1}(\mathbf{x}_2) \\ \phi_{k_2}(\mathbf{x}_1) & \phi_{k_2}(\mathbf{x}_2) \end{pmatrix} = \frac{1}{\sqrt{2}} \left[ \phi_{k_1}(\mathbf{x}_1) \phi_{k_2}(\mathbf{x}_2) - \phi_{k_1}(\mathbf{x}_2) \phi_{k_2}(\mathbf{x}_1) \right], where \phi_k are the single-particle orbitals. The antisymmetry of fermionic basis states manifests as a sign change under odd permutations of particle labels, a direct consequence of the wedge product construction in the algebra. For example, the two-mode state |1_{k_1}, 0_{k_2}\rangle = \hat{a}^\dagger_{k_1} |0\rangle occupies only the first mode, while exchanging modes in a two-particle state |1_{k_1}, 1_{k_2}\rangle = \hat{a}^\dagger_{k_1} \hat{a}^\dagger_{k_2} |0\rangle introduces a minus : \hat{a}^\dagger_{k_2} \hat{a}^\dagger_{k_1} |0\rangle = - |1_{k_1}, 1_{k_2}\rangle. This property distinguishes fermionic Fock states from their bosonic counterparts and is fundamental to describing indistinguishable fermions in quantum many-body systems.

Operator Actions on Fermionic States

In fermionic Fock states, the annihilation operator \hat{a}_k acting on a state with the k-th mode occupied removes the fermion from that mode, yielding \hat{a}_k | \dots 1_k \dots \rangle = | \dots 0_k \dots \rangle, while it annihilates the state if the mode is empty: \hat{a}_k | \dots 0_k \dots \rangle = 0. This binary outcome enforces the Pauli exclusion principle, preventing multiple occupancy in any single mode. The \hat{a}^\dagger_k complements this by adding a to an empty mode: \hat{a}^\dagger_k | \dots 0_k \dots \rangle = | \dots 1_k \dots \rangle, but yields zero when applied to an already occupied mode: \hat{a}^\dagger_k | \dots 1_k \dots \rangle = 0. These actions ensure that occupation numbers n_k remain strictly 0 or 1 across all modes. The number operator \hat{n}_k = \hat{a}^\dagger_k \hat{a}_k measures occupancy in mode k, satisfying \hat{n}_k | \{n\} \rangle = n_k | \{n\} \rangle where n_k \in \{0, 1\}. For an occupied mode, \hat{n}_k | \dots 1_k \dots \rangle = | \dots 1_k \dots \rangle; for empty, \hat{n}_k | \dots 0_k \dots \rangle = 0. Consider a single-mode fermionic Fock state. The vacuum is |0\rangle, and the occupied state is |1\rangle = \hat{a}^\dagger |0\rangle. Here, \hat{a} |1\rangle = |0\rangle, \hat{a} |0\rangle = 0, \hat{a}^\dagger |0\rangle = |1\rangle, and \hat{a}^\dagger |1\rangle = 0, illustrating the exclusion of double occupancy. For a two-mode system with modes k and l, the state |1_k 0_l\rangle = \hat{a}^\dagger_k |0\rangle transforms under \hat{a}_k |1_k 0_l\rangle = |0_k 0_l\rangle and \hat{a}_l |1_k 0_l\rangle = 0, while attempting \hat{a}^\dagger_k |1_k 0_l\rangle = 0 confirms no additional fermion in the occupied mode. Similarly, for |1_k 1_l\rangle = \hat{a}^\dagger_k \hat{a}^\dagger_l |0\rangle, \hat{a}_k |1_k 1_l\rangle = |0_k 1_l\rangle and \hat{a}_l |1_k 1_l\rangle = (-1) |1_k 0_l\rangle, reflecting the antisymmetric nature of the state without allowing further creation in either mode.

General Properties

Multi-Mode Fock States

In , multi-mode Fock states generalize the concept of single-mode Fock states to systems with multiple distinguishable modes, such as different spatial, , or , allowing for a description of particle distributions across these modes while preserving the total particle number. These states form a basis for the in , where each mode is treated as an independent or fermionic degree of freedom. The general form of a multi-mode Fock state for independent modes labeled by k is given by the |\{n_k\}\rangle = \bigotimes_k |n_k\rangle_k, where |n_k\rangle_k denotes the single-mode Fock state with occupation number n_k in mode k. This structure arises naturally in the second-quantized formalism, where for different modes commute (for bosons) or anticommute (for fermions), enabling the construction of multi-particle states without mode mixing. For bosonic particles, the multi-mode Fock state is the direct product of single-mode bosonic number states, with each |n_k\rangle_k = \frac{(\hat{a}_k^\dagger)^{n_k}}{\sqrt{n_k!}} |0\rangle_k, where \hat{a}_k^\dagger is the creation operator for mode k. The total particle number is fixed as N = \sum_k n_k, reflecting the conservation of bosons across modes, and the state is fully symmetric under particle exchange within and across modes due to the bosonic statistics. In the fermionic case, multi-mode Fock states are constructed as antisymmetrized products to enforce the Pauli exclusion principle, with occupation numbers n_k = 0 or $1 per mode. For N fermions occupying distinct modes \alpha_1, \dots, \alpha_N, the state is |\alpha_1, \dots, \alpha_N\rangle = \hat{a}^\dagger_{\alpha_N} \cdots \hat{a}^\dagger_{\alpha_1} |\text{vac}\rangle, up to a normalization factor, and is equivalent to a Slater determinant in the first-quantized representation: \psi(x_1, \dots, x_N) = \frac{1}{\sqrt{N!}} \det \left[ \phi_{\alpha_i}(x_j) \right]_{i,j=1}^N, where \phi_{\alpha_i} are single-particle wavefunctions associated with the modes. This antisymmetric form ensures that interchanging any two fermions introduces a minus sign, capturing the fermionic exchange statistics. Multi-mode Fock states can be represented in various bases, such as or , where the mode indices k correspond to or continuous labels like wavevectors \mathbf{k} or spatial orbitals, yielding states denoted as |n_{\mathbf{k}_1}, n_{\mathbf{k}_2}, \dots \rangle. This flexibility allows adaptation to specific physical systems, such as photonic modes in cavities or orbitals in , while maintaining the occupation-number basis.

Energy Eigenstates and Vacuum Fluctuations

In the context of non-interacting bosonic systems, Fock states serve as energy eigenstates of the free-field Hamiltonian, which takes the form H = \sum_k \omega_k \hat{n}_k in units where \hbar = 1, with \hat{n}_k = \hat{a}_k^\dagger \hat{a}_k denoting the number operator for mode k. The eigenvalue corresponding to a multi-mode Fock state | \{ n_k \} \rangle is E = \sum_k \omega_k n_k, reflecting the definite particle number in each mode and the absence of energy fluctuations within such states. This property arises because the number operators commute with the free Hamiltonian, preserving the occupation numbers under time evolution. However, in interacting systems, such as the Jaynes-Cummings model describing a two-level atom coupled to a single bosonic mode, Fock states are no longer eigenstates; instead, the interaction term mixes bare states into dressed eigenstates with energies split by the coupling strength. A hallmark of quantum fields is the presence of vacuum fluctuations, even in the |0\rangle, where the average field values vanish but uncertainties persist due to non-commuting operators. For the , the operator in a can be expressed as \hat{E} \propto i (\hat{a} e^{-i\omega t} - \hat{a}^\dagger e^{i\omega t}), yielding \langle \hat{E} \rangle = 0 for any Fock state, including the . The variance \Delta E^2 = \langle \hat{E}^2 \rangle - \langle \hat{E} \rangle^2 remains non-zero, specifically \Delta E^2 = \frac{\hbar \omega}{\epsilon_0 V} \left(n + \frac{1}{2}\right) in units for a single of volume V, demonstrating that the (n=0) exhibits a minimum fluctuation level tied to . These fluctuations embody the applied to field quadratures, analogous to position and momentum operators. Defining quadrature operators \hat{X} = (\hat{a} + \hat{a}^\dagger)/\sqrt{2} and \hat{P} = -i (\hat{a} - \hat{a}^\dagger)/\sqrt{2}, which satisfy [\hat{X}, \hat{P}] = i, the vacuum state achieves the minimum uncertainty \Delta X \Delta P = 1/2, with \Delta X = \Delta P = 1/\sqrt{2}. In the position-momentum representation for a single harmonic oscillator mode, the vacuum wavefunction is \psi_0(x) = (\alpha / \pi)^{1/4} e^{-\alpha x^2 / 2}, where \alpha = m \omega / \hbar, leading to \Delta x = \sqrt{\hbar / (2 m \omega)} and \Delta p = \sqrt{m \omega \hbar / 2}, satisfying \Delta x \Delta p = \hbar / 2. This ground-state spreading illustrates how quantum zero-point motion underlies vacuum fluctuations, influencing phenomena like spontaneous emission without requiring excitations.

Non-Classical Behavior

Fock states demonstrate non-classical behavior through their photon number statistics, which deviate markedly from those of classical or coherent light fields. For a pure Fock state |n⟩, the photon number operator â†â has eigenvalue n, yielding a mean photon number ⟨n⟩ = n and zero variance Δn² = 0. This results in sub-Poissonian statistics, where the photon number fluctuations are suppressed below the Poissonian level of coherent states (Δn² = ⟨n⟩). The Mandel Q parameter, defined as Q = (Δn² - ⟨n⟩)/⟨n⟩, evaluates to Q = -1 for any pure Fock state, a negative value that signifies non-classical light with no classical analog. A key manifestation of this non-classicality is , observable in the normalized second-order g^(2)(τ) = ⟨â†(t)â†(t+τ)â(t+τ)â(t)⟩ / ⟨â†â⟩². For the single-photon Fock state |1⟩, g^(2)(0) = 0, indicating that the probability of detecting two photons simultaneously is zero, in stark contrast to classical fields where g^(2)(0) ≥ 1. More generally, for |n⟩ with n ≥ 1, g^(2)(0) = 1 - 1/n < 1, confirming bunching suppression that strengthens with smaller n. Fock states also exhibit non-classical features in their projections, with variances ΔX² = ΔP² = n + 1/2 (using the definitions above), larger than the value of 1/2 and reflecting the complete delocalization of due to the fixed particle number. This symmetry in quadrature uncertainties, combined with the increased product ΔX ΔP = n + 1/2, underscores their departure from classical wave descriptions, where both and are well-defined without such extremes. Unlike coherent states, which undergo due to photon number fluctuations, Fock states experience no such , as their definite particle number eliminates dephasing mechanisms tied to number-dependent evolutions.

Applications

Single-Photon Sources

The single-photon Fock state, denoted as |1\rangle, represents an ideal single-photon state in , characterized by exactly one in a given with no probability of (|0\rangle) or multi-photon components. This purity ensures deterministic photon number, distinguishing it from coherent or sources that exhibit Poissonian statistics and potential multi-photon emissions. Generation of |1\rangle states often relies on heralded sources using (SPDC), where a photon splits into correlated signal and idler pairs in a nonlinear ; detection of the idler heralds the presence of the signal in the |1\rangle state. Quantum dots provide an alternative deterministic approach, where resonant excitation of a single leads to radiative decay emitting a in the |1\rangle Fock state, achieving high indistinguishability and on-demand operation. In (QKD), such as the protocol, |1\rangle states enable secure single-photon transmission by preventing eavesdropping attacks that rely on multi-photon vulnerabilities in attenuated sources. The exact photon number ensures , as any interception disturbs the detectably. Despite these advances, generating pure |1\rangle states faces challenges, including imperfect heralding efficiency in SPDC due to multi-mode emissions and losses, which result in mixed states rather than pure Fock states. sources, while brighter, suffer from residual multi-photon events from background excitations or re-excitation, achieving very high single-photon purity, with multi-photon emission probabilities suppressed to as low as 10^{-4}, although residual events from background excitations or re-excitation can occur in some configurations.

Quantum Information Contexts

In continuous-variable , bosonic Fock states serve as essential non-Gaussian resources for qumodes, where the photon number n in the state |n\rangle encodes logical information, enabling operations beyond the limitations of Gaussian states alone. These states allow for the implementation of universal quantum gates through measurements and feed-forward, addressing the inability of purely Gaussian resources to generate entanglement in certain protocols. For instance, injecting Fock states into optical modes facilitates the of complex Hamiltonians and error-corrected computation in photonic or superconducting platforms. Fermionic Fock states play a central role in second-quantized simulations of many-body systems on , representing exact occupation numbers in orbital bases that are mapped to configurations via compact encodings. This approach enables efficient Trotterization of fermionic Hamiltonians, such as those in , by preserving antisymmetry and avoiding sign problems inherent in classical simulations. Recent implementations on superconducting processors have demonstrated the preparation and manipulation of these states for modeling electronic structures with reduced overhead. Multi-mode Fock states, consisting of definite particle distributions across multiple bosonic modes, underpin protocols like on photonic chips, where single-photon Fock states in distinct input modes interfere via linear optics to produce output distributions hard to simulate classically. These states can be combined with Gaussian operations, such as squeezing, to generate hybrid resources for sampling tasks that approximate or extend traditional Fock-based sampling, enhancing scalability in integrated . The precise control of particle number in Fock states provides key advantages in processing, including inherent detection of errors like photon loss or gain through number-resolved measurements, which supports error-resistant and fault-tolerant encoding. In bosonic error correction codes, superpositions involving squeezed Fock states offer exponential suppression of and loss errors, outperforming cat codes in certain regimes. Post-2020 advances in superconducting circuits have realized Fock-state-based bosonic qubits with times exceeding one , enabling modular architectures for scalable quantum processors. In quantum metrology, large Fock states have enabled Heisenberg-limited phase estimation, with experiments demonstrating enhanced sensitivity using up to 40 as of 2024.

References

  1. [1]
    None
    ### Definition and Properties of Fock States
  2. [2]
  3. [3]
    Quantum simulations with multiphoton Fock states - Nature
    Jun 3, 2021 · Here we show a quantum simulation performed using genuine higher-order Fock states, with two or more indistinguishable particles occupying the same bosonic ...
  4. [4]
    [PDF] Generation of Fock states in a superconducting quantum circuit
    Quantum behaviour in an oscillator is most obvious in Fock states, states with specific numbers of energy quanta, but such states are hard to create3–7. Here we ...Missing: mechanics | Show results with:mechanics
  5. [5]
    [PDF] Quantum Mechanics
    Second quantization is a formalism used to describe quantum many-body systems of iden- tical particles.  Classical mechanics: each particle is labeled by a ...
  6. [6]
  7. [7]
    [PDF] 1. Second Quantization - Giuseppe De Nittis
    May 1, 2016 · Second quantization. Vladimir Fock (1898–1974). Vladimir Fock received his education from the University of Petrograd in difficult times. In ...
  8. [8]
    [PDF] Selected Works V.A.Fock
    In those works, Fock establishes the coherent theory of second quantization introducing the Fock space, puts forward the Fock method of functionals, introduces ...
  9. [9]
    [PDF] Creation and Annihilation Operators Occupation number ...
    A feature of the Fock space is that the total number of particles is not a fixed parameter, but rather is a dynamical variable associated with a total number ...
  10. [10]
    [PDF] An Introduction to Second Quantization
    Jan 27, 2010 · Since, we are interested in situations where particle number is not fixed, we want operators which can take us from one Fock space to another ...
  11. [11]
    [PDF] Creation and Annihilation Operators - CMU Quantum Theory Group
    It spans a. one-dimensional subspace HS. 0 of the Fock space, and is annihilated by every one of the annihilation operators: bj|∅i = 0. (6)
  12. [12]
    [PDF] arXiv:1111.0787v3 [math-ph] 1 Jun 2012
    Jun 1, 2012 · because of the Fock vacuum state, which is a unique joint eigenstate ... and annihilation operator smeared with functions well localized in ...
  13. [13]
    [PDF] Today: combining SR and QM
    many particles/antiparticles of any type are present, even in the lowest energy state, i.e. the vacuum. • The lowest energy state is not one in which we can say ...
  14. [14]
    The quantum theory of the emission and absorption of radiation
    The quantum theory of the emission and absorption of radiation. Paul Adrien Maurice Dirac.
  15. [15]
    [PDF] “Second quantization” (the occupation-number representation)
    Feb 14, 2013 · The second quantization method involves the use of so-called creation and annihilation oper- ators. These operators respectively create and ...
  16. [16]
    [PDF] Pseudo-Boson Coherent and Fock States - arXiv
    Feb 21, 2009 · Fock state wave functions are obtained as product of an exponential of a quadratic form of x and one of the two new polynomials Pn(x), Qn(x) ...
  17. [17]
    Pair-correlation ansatz for the ground state of interacting bosons in ...
    Jun 23, 2021 · Efficient construction of many-body Fock states having the lowest energies. Acta Phys. Pol. A 136, 566–570. https://doi.org/10.12693 ...Formulation Of The Problem · Variational Ansatz · Two-Particle Case
  18. [18]
    [PDF] arXiv:2211.05714v1 [quant-ph] 10 Nov 2022
    Nov 10, 2022 · I review bosonic quantum memories, characterizing them as either bosonic stabilizer or bosonic Fock-state codes. I then enumerate various ...<|control11|><|separator|>
  19. [19]
    [1805.04552] Back and forth from Fock space to Hilbert space - arXiv
    May 11, 2018 · Abstract page for arXiv paper 1805.04552: Back and forth from Fock space to Hilbert space: a guide for commuters.
  20. [20]
    [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,. 2003). MH. G. D. Mahan, Many-Particle Physics (Kluwer ...
  21. [21]
    [PDF] Fermionic Algebra and Fock Space - UT Physics
    it is killed by all the annihilation operators вp,s and bp,s in the set. The two types of cre- ation operators в.
  22. [22]
    [PDF] 42 Fermion annihilation and creation operators - David Miller
    Anticommutation relation for mixed operators. Hence we can say that. Putting this together with we can write the anticommutation relation for mixed annihilation ...
  23. [23]
    [PDF] 1 Definition of creation and annihilation operators
    we get the anticommutation relation: (c. † k′ , c. † k)+ = 0. 1. Page 2. Quantum Mechanics II. Lecture 36. In general, we get the following anticommutation ...
  24. [24]
    [PDF] Untitled - Scalettar
    creation and annihilation operators obey the anticommutation rela- tions. {ao, ao t} = 1. {a₁, a₁t} - 1. SECOND QUANTIZATION. 417. Ino, n₁, ng,. • .) (az tyng.
  25. [25]
    [PDF] Chapter 4 Introduction to many-body quantum mechanics
    Note that while the set of Slater determinants of single particle basis functions forms a basis of the fermionic Fock space, the general fermionic many body ...
  26. [26]
    [PDF] Fermionic Quasi-free States and Maps in Information Theory - arXiv
    The first term in the direct sum is the vacuum state and the last one is the completely filled Fermi sea.
  27. [27]
  28. [28]
  29. [29]
    [PDF] Review of Quantum Electromagnetic States - DTIC
    May 5, 2014 · 4.2.2 Fock States as Energy Eigenstates. 4.2.3 Fluctuations of EM Fields in Fock States. 4.3 Introduction to EM Coherent States. 4.3.1 The ...
  30. [30]
    [PDF] Comparison of Quantum and Semiclassical Radiation Theory with ...
    1963. Jaynes and Cummings: Quantum and Semiclassical Radiation Theories tions to find the transient response to an arbitrary small perturbation. Thus ...
  31. [31]
    [PDF] arXiv:2310.04913v1 [quant-ph] 7 Oct 2023
    Oct 7, 2023 · This can be quantified via Mandel's Q parameter, defined as Q = (∆n)2 / <n> - 1; it has a minimum value of Q = -1 for Fock states and is null ...
  32. [32]
    Transforming photon statistics through zero-photon subtraction
    Apr 25, 2023 · Experimentally, these properties can be conveniently characterized by Mandel's Q parameter, with Q > 0 for super-Poissonian sources and − 1 ≤ Q ...
  33. [33]
    [PDF] Lecture 4: Quantum states of light — Fock states • Definition
    The fluctuations must vanish because the value of the photon number is 'sharp' in a Fock state. For the same reason, there are no energy fluctuations in a Fock ...
  34. [34]
    Nonclassical Light - RP Photonics
    Fock states* (also called photon number states) have a well-defined number of photons, whereas the optical phase is totally undefined. · Squeezed states of light ...Missing: diffusion | Show results with:diffusion
  35. [35]
    Controlling Photon Number Coherence for Quantum Cryptography
    Jan 27, 2024 · In the ideal case, a single photon can be represented by the one-photon Fock state, but due to imperfections or coupling to the environment, ...
  36. [36]
    Single-photon sources: Approaching the ideal through multiplexing
    Apr 30, 2020 · Sources with the Schmidt number greater than 1 ... Boson sampling with single-photon fock states from a bright solid-state source. ,”.Single-photon sources · Single photons: Definitions · Probabilistic single-photon...
  37. [37]
    Optimized generation of heralded Fock states using parametric ...
    The generation of heralded pure Fock states via spontaneous parametric down-conversion (PDC) relies on perfect photon-number correlations in the output modes.
  38. [38]
    Single-photon Sources – on demand, heralded, multiplexed ...
    The ideal result is a one-photon Fock state in a well-defined spatial–temporal mode, but practically, purity, brightness and indistinguishability (see below) ...
  39. [39]
    Quantum dot single-photon sources with ultra-low multi ... - Nature
    Sep 14, 2018 · Loredo, J. C. et al. Boson sampling with single-photon fock states from a bright solid-state source. Phys. Rev. Lett. 118, 130503 (2017).
  40. [40]
    [PDF] Quantum simulation of second-quantized Hamiltonians in compact ...
    Oct 13, 2021 · In this section, we define the compact encoding, which maps Fock states to qubit states, and the input model for our second-quantized ...
  41. [41]
    Boson sampling with Gaussian measurements | Phys. Rev. A
    Sep 14, 2017 · ... state of the modes on B side. This partially time-reversed or time-unfolded TSBS results in a boson sampling setting with a multimode Fock state ...
  42. [42]
    Reconfigurable continuously-coupled 3D photonic circuit for Boson ...
    May 12, 2022 · In the GBS paradigm, Fock states are replaced by single-mode squeezed vacuum states (see Fig. ... Boson sampling from a gaussian state.Abstract · Results · Unitary Matrix Sampling And...
  43. [43]
    Surpassing millisecond coherence in on chip superconducting ...
    May 1, 2024 · These results demonstrate an advancement towards a more modular and compact coaxial circuit architecture for bosonic qubits with reproducibly ...Results · Methods · Transmon Coupling Quality...