Fact-checked by Grok 2 weeks ago

Quantization of the electromagnetic field

The quantization of the is a foundational concept in quantum physics that applies the principles of quantum mechanics to the classical electromagnetic field described by , treating it as a system of infinite harmonic oscillators whose energy is quantized into discrete units called photons. This process transforms the continuous classical field variables—such as the E and B—into non-commuting operators, enabling the description of as a stream of massless spin-1 bosons with energy ℏω, where is the reduced Planck's constant and ω is the . The resulting theory, known as (QED), accounts for interactions between light and matter at the quantum level, including vacuum fluctuations and . The development of electromagnetic field quantization emerged in the mid-1920s amid efforts to reconcile with and . Pioneering work began with Pascual Jordan's 1925 quantization of the free using , followed by Paul Dirac's 1927 formulation of , which introduced the idea of the field as an operator to describe radiation emission and absorption by atoms. Contributions from , , and in 1926–1929 further solidified the procedure, treating the field as an infinite collection of analogous to mechanical oscillators. This "second quantization" approach, as it became known, resolved inconsistencies in early quantum theories of radiation and laid the groundwork for modern . In the standard quantization procedure, the is confined to a finite volume, such as a cubic with , to discretize the modes. The A(r, t) is expanded in a of plane-wave normal modes u_{kλ}(r), where k denotes the wave vector and λ the , yielding coefficients that serve as q_{kλ} and momenta p_{kλ}. These are promoted to operators satisfying commutation relations [Q_{kλ}, P_{k'λ'}] = iℏ δ_{kk'} δ_{λλ'}, and further expressed via boson creation a†_{kλ} and a_{kλ} operators with [a_{kλ}, a†{k'λ'}] = δ{kk'} δ_{λλ'}. The then takes the form H = ∑{kλ} ℏω_k (a†{kλ} a_{kλ} + 1/2), revealing discrete energy levels and a non-zero () even in the . Gauge choices, such as the or gauge, ensure transversality (∇ · A = 0) and compatibility with . This quantization framework underpins , the paradigmatic , and has profound implications for , enabling predictions of phenomena like the —arising from vacuum fluctuations between conducting plates, first theorized by Hendrik in 1948 and experimentally confirmed in 2001. It also facilitates the study of photon statistics in Fock states |n_{kλ}⟩, light-matter interactions, and extensions to media with dielectrics or magnetic monopoles. Modern applications include processing and precision measurements, underscoring its enduring role in fundamental physics.

Classical Electromagnetic Field

Vector Potential and Fields

In the mid-19th century, James Clerk Maxwell introduced the concept of electromagnetic potentials as part of his dynamical theory of the , building on earlier work by and William Thomson. Maxwell's seminal 1865 paper utilized a quantity he termed "electromagnetic momentum," which corresponds to the modern , to express the relations between electric and magnetic forces. The was mathematically formalized by Thomson in 1851 as a means to represent the , providing a foundation that Maxwell later incorporated into his equations. The is a relativistic A^\mu = (\phi/c, \mathbf{A}), where \phi is the scalar , \mathbf{A} is the , and c is the in vacuum. This formulation combines the scalar and vector potentials into a single entity that transforms covariantly under Lorentz transformations, unifying the description of electromagnetic phenomena in . The four-potential satisfies the inhomogeneous \partial_\nu F^{\mu\nu} = \mu_0 J^\mu, where F^{\mu\nu} = \partial^\mu A^\nu - \partial^\nu A^\mu is the electromagnetic field tensor and J^\mu is the four-current. The electric and magnetic fields can be derived from the four-potential components. In three-vector notation, the electric field is given by \mathbf{E} = -\nabla \phi - \frac{\partial \mathbf{A}}{\partial t}, which arises from the antisymmetric nature of the field tensor F^{\mu\nu}, capturing both the conservative part from the scalar potential gradient and the induced part from the time-varying vector potential. The magnetic field is \mathbf{B} = \nabla \times \mathbf{A}, obtained directly as the curl of the vector potential, ensuring the divergence-free condition \nabla \cdot \mathbf{B} = 0 is automatically satisfied. These expressions demonstrate how the potentials encode the observable fields while introducing redundancy due to gauge invariance. The choice of potentials is not unique, reflecting the gauge freedom inherent in . A gauge transformation A^\mu \to A^\mu + \partial^\mu \Lambda for an arbitrary scalar \Lambda leaves the field tensor F^{\mu\nu} unchanged, thus preserving the physical fields \mathbf{E} and \mathbf{B}. Common gauge choices include the , \partial_\mu A^\mu = 0, which simplifies the wave equations for the potentials to the form \Box A^\mu = -\mu_0 J^\mu and is Lorentz invariant. Another is the , \nabla \cdot \mathbf{A} = 0, which decouples the scalar potential to satisfy instantaneously and is useful in non-relativistic contexts or for problems.

Lagrangian and Hamiltonian in Terms of Potentials

The density for the free in is given by \mathcal{L} = -\frac{1}{4} F_{\mu\nu} F^{\mu\nu}, where the electromagnetic field strength tensor is defined as F_{\mu\nu} = \partial_\mu A_\nu - \partial_\nu A_\mu, with A^\mu = (\phi, \mathbf{A}) denoting the four-potential comprising the \phi and the \mathbf{A}. This form ensures Lorentz invariance and gauge invariance under transformations A^\mu \to A^\mu + \partial^\mu \Lambda. In three-vector notation and using Heaviside-Lorentz units (with c = 1), the density simplifies to \mathcal{L} = \frac{1}{2} (\mathbf{E}^2 - \mathbf{B}^2), where the electric field is \mathbf{E} = -\nabla \phi - \partial_t \mathbf{A} and the magnetic field is \mathbf{B} = \nabla \times \mathbf{A}. The action functional S = \int \mathcal{L} \, d^4x is stationary under variations of the potentials, leading via the Euler-Lagrange equations \partial_\mu \left( \frac{\partial \mathcal{L}}{\partial (\partial_\mu A_\nu)} \right) - \frac{\partial \mathcal{L}}{\partial A_\nu} = 0 to the homogeneous and inhomogeneous Maxwell equations in vacuum: \partial^\mu F_{\mu\nu} = 0 and \partial_{[\lambda} F_{\mu\nu]} = 0. To transition to the Hamiltonian formulation, the canonical momentum conjugate to the vector potential \mathbf{A} is computed as \boldsymbol{\Pi} = \frac{\partial \mathcal{L}}{\partial (\partial_t \mathbf{A})} = -\mathbf{E}, while the scalar potential \phi has no conjugate momentum, rendering it non-dynamical. The Hamiltonian density then follows from the Legendre transform \mathcal{H} = \boldsymbol{\Pi} \cdot \partial_t \mathbf{A} - \mathcal{L}, yielding \mathcal{H} = \frac{1}{2} (\boldsymbol{\Pi}^2 + \mathbf{B}^2). This expression represents the total of the , with the first term corresponding to the contribution and the second to the magnetic. In the Hamiltonian framework, Hamilton's equations generate the field dynamics, but the absence of a conjugate to \phi implies that the \nabla \cdot \mathbf{E} = 0 ( in free space) emerges as a primary rather than a dynamical . This must be preserved under , enforcing consistency with the transverse nature of in vacuum.

Quantization Procedure

Gauge Choice and Mode Expansion

In the quantization of the electromagnetic field, particularly in non-relativistic contexts, the Coulomb gauge is selected, defined by the \nabla \cdot \mathbf{A} = 0. This choice is preferred because it eliminates the unphysical longitudinal modes from the dynamical quantization, restricting the quantum to the two transverse polarizations that correspond to physical . In this gauge, the longitudinal arises instantaneously from the via the \phi = -\int \frac{\rho(\mathbf{r}')}{4\pi\epsilon_0 |\mathbf{r} - \mathbf{r}'|} d^3\mathbf{r}', treated classically rather than as a propagating quantum mode, simplifying the structure for matter-field interactions. The vector potential \mathbf{A} in the Coulomb gauge is decomposed into a Fourier mode expansion suitable for the free field, providing a basis of plane waves that diagonalize the classical wave equation. In free space, this takes the form \mathbf{A}(\mathbf{r}, t) = \int \frac{d^3k}{(2\pi)^3} \sum_{\lambda=1,2} \left[ a_{\mathbf{k},\lambda} \boldsymbol{\epsilon}_{\mathbf{k},\lambda} e^{i(\mathbf{k}\cdot\mathbf{r} - \omega t)} + \mathrm{h.c.} \right], where \omega = c |\mathbf{k}| relates the angular frequency to the wave vector magnitude, the coefficients a_{\mathbf{k},\lambda} are complex amplitudes (to be promoted to annihilation operators in quantization), and \boldsymbol{\epsilon}_{\mathbf{k},\lambda} (\lambda = 1, 2) are unit polarization vectors. This expansion originates from the classical Hamiltonian expressed in terms of \mathbf{A} and its conjugate momentum, facilitating the identification of normal modes analogous to harmonic oscillators. The polarization vectors \boldsymbol{\epsilon}_{\mathbf{k},\lambda} satisfy the \mathbf{k} \cdot \boldsymbol{\epsilon}_{\mathbf{k},\lambda} = 0, ensuring the and resulting fields \mathbf{E} = -\partial_t \mathbf{A} and \mathbf{B} = \nabla \times \mathbf{A} have no longitudinal components, with the two orthogonal polarizations spanning the plane perpendicular to \mathbf{k} such that \boldsymbol{\epsilon}_{\mathbf{k},\lambda} \cdot \boldsymbol{\epsilon}_{\mathbf{k},\lambda'}^* = \delta_{\lambda\lambda'}. This condition enforces the physical content of in vacuum, where only transverse waves propagate. The form of the mode expansion depends on boundary conditions: in unbounded free space, the continuous of \mathbf{k} yields the volume over d^3k, normalized by (2\pi)^3 for delta-function of plane waves. In a finite , such as a cubic box of volume V = L^3 with , the modes discretize to \mathbf{k} = \frac{2\pi}{L} (n_x, n_y, n_z) with integers n_i, replacing the with a sum \frac{1}{V} \sum_{\mathbf{k}}, which converges to the continuous limit as V \to \infty.

Canonical Quantization and Commutation Relations

In the approach to the , the classical \mathbf{A}(\mathbf{r}, t) and its conjugate momentum density \boldsymbol{\Pi}(\mathbf{r}, t) = -\epsilon_0 \mathbf{E}(\mathbf{r}, t), derived from the , are promoted to non-commuting operators \hat{\mathbf{A}}(\mathbf{r}, t) and \hat{\boldsymbol{\Pi}}(\mathbf{r}, t). This promotion, first systematically applied to the radiation field by Dirac, transforms the into a quantum mechanical framework by imposing on the dynamical variables. The fundamental equal-time commutation relations for these operators are [\hat{A}_i(\mathbf{r}, t), \hat{\Pi}_j(\mathbf{r}', t)] = i \hbar \delta_{ij} \delta^3(\mathbf{r} - \mathbf{r}'), with all other commutators among components of \hat{\mathbf{A}} and \hat{\boldsymbol{\Pi}} vanishing at equal times. These relations ensure the consistency of the , analogous to the position-momentum in single-particle , and follow directly from the canonical structure of the field theory. Building on the mode expansion of the classical field in a chosen gauge, the operator \hat{\mathbf{A}}(\mathbf{r}, t) is expressed in terms of creation \hat{a}^\dagger_{\mathbf{k},\lambda} and annihilation \hat{a}_{\mathbf{k},\lambda} operators as \hat{\mathbf{A}}(\mathbf{r}, t) \propto \sum_{\mathbf{k},\lambda} \left( \hat{a}_{\mathbf{k},\lambda} \boldsymbol{\epsilon}_{\mathbf{k},\lambda} \, e^{i \mathbf{k} \cdot \mathbf{r} - i \omega_k t} + \hat{a}^\dagger_{\mathbf{k},\lambda} \boldsymbol{\epsilon}^*_{\mathbf{k},\lambda} \, e^{-i \mathbf{k} \cdot \mathbf{r} + i \omega_k t} \right), where \boldsymbol{\epsilon}_{\mathbf{k},\lambda} denotes the polarization vector for mode \mathbf{k} and helicity \lambda, and the proportionality includes normalization factors involving \hbar, \epsilon_0, \omega_k = c |\mathbf{k}|, and the quantization volume. The creation and annihilation operators obey the bosonic commutation algebra [\hat{a}_{\mathbf{k},\lambda}, \hat{a}^\dagger_{\mathbf{k}',\lambda'}] = \delta_{\mathbf{k},\mathbf{k}'} \delta_{\lambda,\lambda'}, with all other commutators zero, which is derived from the canonical relations to preserve the quantum structure. This framework operates in the , where the field operators evolve according to the classical field equations while acting on a fixed of states, distinguishing it from first-quantized single-particle . In the context of , the field operators directly create and annihilate particles (photons) in the , providing a many-body description essential for .

Quantized Field Hamiltonian

Expansion in Normal Modes

Following gauge fixing to the Coulomb gauge, where the scalar potential is set to zero and the vector potential \mathbf{A} is transverse (\nabla \cdot \mathbf{A} = 0), the constraints are satisfied, and the Hamiltonian for the free electromagnetic field takes the form \hat{H} = \int d^3 r \left[ \frac{1}{2} \hat{\boldsymbol{\Pi}}^2(\mathbf{r}) + \frac{1}{2} (\nabla \times \hat{\mathbf{A}}(\mathbf{r}))^2 \right], with \hat{\boldsymbol{\Pi}}(\mathbf{r}) = -\dot{\hat{\mathbf{A}}}(\mathbf{r}) serving as the canonical momentum conjugate to \hat{\mathbf{A}}, in units where \epsilon_0 = \mu_0 = 1. The vector potential is expanded in terms of transverse plane-wave normal modes as \hat{\mathbf{A}}(\mathbf{r}) = \sum_{\mathbf{k},\lambda} \sqrt{\frac{\hbar}{2 \omega_{\mathbf{k}} V}} \left[ \hat{a}_{\mathbf{k},\lambda} \boldsymbol{\epsilon}_{\mathbf{k},\lambda} e^{i \mathbf{k} \cdot \mathbf{r}} + \hat{a}^\dagger_{\mathbf{k},\lambda} \boldsymbol{\epsilon}^*_{\mathbf{k},\lambda} e^{-i \mathbf{k} \cdot \mathbf{r}} \right], where \mathbf{k} labels the wave vector, \lambda = 1,2 denotes the two transverse polarization states with unit vectors \boldsymbol{\epsilon}_{\mathbf{k},\lambda}, \omega_{\mathbf{k}} = c |\mathbf{k}| is the mode frequency, V is the quantization volume, and \hat{a}_{\mathbf{k},\lambda}, \hat{a}^\dagger_{\mathbf{k},\lambda} are the annihilation and creation operators satisfying bosonic commutation relations from canonical quantization. Substituting this mode expansion into the Hamiltonian, along with the corresponding expansion for \hat{\boldsymbol{\Pi}}(\mathbf{r}), yields a diagonal form \hat{H} = \sum_{\mathbf{k},\lambda} \hbar \omega_{\mathbf{k}} \left( \hat{a}^\dagger_{\mathbf{k},\lambda} \hat{a}_{\mathbf{k},\lambda} + \frac{1}{2} \right), where the orthogonality of the modes ensures no cross terms between different \mathbf{k} and \lambda. The term \sum_{\mathbf{k},\lambda} \frac{1}{2} \hbar \omega_{\mathbf{k}} represents the zero-point energy of the vacuum, which diverges due to the infinite number of modes in continuous space. This infinity is regularized through normal ordering, redefining the Hamiltonian as \hat{H} = \sum_{\mathbf{k},\lambda} \hbar \omega_{\mathbf{k}} \hat{a}^\dagger_{\mathbf{k},\lambda} \hat{a}_{\mathbf{k},\lambda}, which sets the ground-state energy to zero while preserving energy differences observable in phenomena like the Casimir effect. In the , known as the , the expectation value of the field operators vanishes (\langle 0 | \hat{\mathbf{A}}(\mathbf{r}) | 0 \rangle = 0), yet vacuum fluctuations persist, manifesting as non-zero variances \langle 0 | \hat{\mathbf{A}}^2(\mathbf{r}) | 0 \rangle > 0 that arise from the commutation relations and contribute to effects such as and the .

Analogy to Quantum Harmonic Oscillators

In the classical description of the free , the satisfies the wave equation derived from in , allowing a decomposition into plane-wave normal modes labeled by \mathbf{k} and index \lambda. These modes represent independent transverse oscillations, decoupled in Fourier space, where each behaves dynamically as a classical with \omega_k = c k (with k = |\mathbf{k}| and c the ). Upon canonical quantization, this structure maps directly to quantum mechanics, treating each mode (\mathbf{k}, \lambda) as an independent one-dimensional quantum harmonic oscillator. The canonical variables of the mode—the "position" coordinate q_{\mathbf{k}\lambda} (proportional to the mode amplitude of the vector potential) and conjugate "momentum" p_{\mathbf{k}\lambda} (proportional to the mode's electric field component)—become operators satisfying [q_{\mathbf{k}\lambda}, p_{\mathbf{k}'\lambda'}] = i\hbar \delta_{\mathbf{k}\mathbf{k}'} \delta_{\lambda\lambda'}. These are expressed in terms of creation \hat{a}^\dagger_{\mathbf{k}\lambda} and annihilation \hat{a}_{\mathbf{k}\lambda} operators via q_{\mathbf{k}\lambda} \propto \sqrt{\frac{\hbar}{2\omega_k}} \left( \hat{a}_{\mathbf{k}\lambda} + \hat{a}^\dagger_{\mathbf{k}\lambda} \right), \quad p_{\mathbf{k}\lambda} \propto i \sqrt{\frac{\hbar \omega_k}{2}} \left( \hat{a}^\dagger_{\mathbf{k}\lambda} - \hat{a}_{\mathbf{k}\lambda} \right), with the commutation relations [\hat{a}_{\mathbf{k}\lambda}, \hat{a}^\dagger_{\mathbf{k}'\lambda'}] = \delta_{\mathbf{k}\mathbf{k}'} \delta_{\lambda\lambda'}. The resulting Hamiltonian for each mode mirrors that of a quantum harmonic oscillator, summing over all modes to yield the total quantized field energy. The energy spectrum for each mode consists of discrete levels E_n = \hbar \omega_k \left( n + \frac{1}{2} \right), where n = 0, 1, 2, \dots labels the excitation number, introducing the zero-point energy \frac{1}{2} \hbar \omega_k even in the ground state. This quantization framework provides intuitive insight into the field's particle-like excitations (photons) while preserving its wavelike mode structure. The analogy originated in Paul Dirac's seminal 1927 paper, which first applied quantum rules to the radiation field by treating its monochromatic components as quantized oscillators to reconcile emission and absorption processes with quantum mechanics.

Photon States

Fock States and Photon Number Operator

In the quantized electromagnetic field, the number operator for a specific labeled by \mathbf{k} and \lambda is defined as \hat{N}_{\mathbf{k},\lambda} = \hat{a}^\dagger_{\mathbf{k},\lambda} \hat{a}_{\mathbf{k},\lambda}, where \hat{a}_{\mathbf{k},\lambda} and \hat{a}^\dagger_{\mathbf{k},\lambda} are the annihilation and creation operators, respectively. This operator has eigenvalues n_{\mathbf{k},\lambda} = 0, 1, 2, \dots , corresponding to the number of occupying that , reflecting the discrete nature of the field's energy quanta. The eigenstates of the total number operator \hat{N} = \sum_{\mathbf{k},\lambda} \hat{N}_{\mathbf{k},\lambda} form the for the field, constructed as multi-mode tensor products. A general is expressed as | \{ n_{\mathbf{k},\lambda} \} \rangle = \prod_{\mathbf{k},\lambda} \frac{ (\hat{a}^\dagger_{\mathbf{k},\lambda})^{n_{\mathbf{k},\lambda}} }{\sqrt{n_{\mathbf{k},\lambda}!}} |0\rangle, where the product runs over all modes, and |0\rangle denotes the vacuum state. These states provide a complete orthonormal basis for the of the field, with fixed total number N = \sum_{\mathbf{k},\lambda} n_{\mathbf{k},\lambda}, analogous to the energy eigenstates of a system of independent harmonic oscillators. The state |0\rangle is uniquely defined as the state annihilated by all annihilation operators, satisfying \hat{a}_{\mathbf{k},\lambda} |0\rangle = 0 for every mode \mathbf{k},\lambda. Despite containing zero photons, this state exhibits non-zero expectation values for the squares of the electric and magnetic field operators, \langle 0| \hat{\mathbf{E}}^2 |0 \rangle > 0 and \langle 0| \hat{\mathbf{B}}^2 |0 \rangle > 0, arising from the infinite sum over zero-point fluctuations across all modes. These vacuum fluctuations underscore the quantum nature of the even in the absence of excitations. The bosonic character of photons stems from the commutation relations [\hat{a}_{\mathbf{k},\lambda}, \hat{a}^\dagger_{\mathbf{k}',\lambda'}] = \delta_{\mathbf{k},\mathbf{k}'} \delta_{\lambda,\lambda'}, which enforce symmetric statistics under particle exchange. This leads to indistinguishable photons within the same mode, consistent with the foundational treatment of radiation quanta as bosons obeying Bose-Einstein statistics.

Coherent States (brief introduction)

In , coherent states provide a quantum description of that closely resembles classical electromagnetic , particularly those produced by lasers, where the field exhibits stable and . These states are constructed as superpositions of Fock states, the number eigenstates from the basis, and are defined as the eigenstates of the annihilation operator \hat{a}, satisfying \hat{a} |\alpha\rangle = \alpha |\alpha\rangle, where \alpha is a eigenvalue. Explicitly, the coherent state is given by |\alpha\rangle = e^{-|\alpha|^2/2} \sum_{n=0}^\infty \frac{\alpha^n}{\sqrt{n!}} |n\rangle, which normalizes the state and weights the Fock components |n\rangle according to a . The photon number operator \hat{N} = \hat{a}^\dagger \hat{a} in a coherent state yields an expectation value \langle \hat{N} \rangle = |\alpha|^2, with the variance equal to the mean, \Delta N^2 = |\alpha|^2, characteristic of the Poissonian statistics that minimize quantum noise relative to the intensity. Under free evolution governed by the harmonic oscillator Hamiltonian \hat{H} = \hbar \omega (\hat{N} + 1/2), the coherent state parameter evolves as \alpha(t) = |\alpha| e^{-i \omega t}, preserving the shape of the wave packet while rotating its phase in the complex plane. Coherent states, introduced by in 1963, form the foundation of modern , enabling the analysis of coherent phenomena like and superposition in the quantized .

Photon Properties

Energy and Frequency Relation

In the quantized , the energy of a single associated with a plane-wave mode of \mathbf{k} and \lambda is E = \hbar \omega_k, where \omega_k = c |\mathbf{k}| is the and c is the in . This relation follows from the for electromagnetic in free space, \omega_k = c k, yielding the equivalent form E = \hbar c k. Equivalently, in terms of the frequency f = \omega_k / 2\pi, the energy is E = h f, where h = 2\pi \hbar is Planck's constant. This energy-frequency duality was first proposed by in 1900 as part of his derivation of the spectrum, where he postulated that the energy of electromagnetic oscillators is quantized in discrete units to resolve the predicted by classical Rayleigh-Jeans theory. Planck's hypothesis, formalized in his 1901 paper, introduced the constant h such that the average energy of a resonator at frequency \nu (equivalent to f) is \langle E \rangle = h \nu / (e^{h \nu / kT} - 1), leading to the spectral energy density that matched experimental observations. Although Planck initially viewed quantization as a mathematical trick for classical oscillators rather than a fundamental property of radiation, this relation laid the groundwork for the concept later developed by Einstein in 1905. For multi-photon states in the free field, the total is additive across modes due to the non-interacting nature of . In the Fock basis, the acting on a |\{n_{\mathbf{k},\lambda}\}\rangle with occupation numbers n_{\mathbf{k},\lambda} yields \hat{H} |\{n\}\rangle = \left( \sum_{\mathbf{k},\lambda} \hbar \omega_k n_{\mathbf{k},\lambda} \right) |\{n\}\rangle, excluding the infinite \sum_{\mathbf{k},\lambda} \frac{1}{2} \hbar \omega_k which is typically renormalized away in free-field theory. This reflects the bosonic statistics of , allowing arbitrary occupation numbers per while preserving as a linear sum over individual contributions.

Momentum and Wave Vector

In the quantized description of the , the total linear operator is expressed as \hat{\mathbf{P}} = -\int d^3 r \, \hat{\mathbf{E}}(\mathbf{r}) \times \hat{\mathbf{B}}(\mathbf{r}), where \hat{\mathbf{E}}(\mathbf{r}) and \hat{\mathbf{B}}(\mathbf{r}) are the quantum electric and operators evaluated at position \mathbf{r}. This form arises from the classical electromagnetic density, adapted to the quantum regime. Expanding the field operators in the normal mode basis of plane waves with wave vector \mathbf{k} and polarization \lambda, the momentum operator simplifies to \hat{\mathbf{P}} = \sum_{\mathbf{k},\lambda} \hbar \mathbf{k} \, \hat{a}^\dagger_{\mathbf{k},\lambda} \hat{a}_{\mathbf{k},\lambda}, where \hat{a}^\dagger_{\mathbf{k},\lambda} and \hat{a}_{\mathbf{k},\lambda} are the creation and annihilation operators for photons in mode (\mathbf{k}, \lambda). The number operator \hat{n}_{\mathbf{k},\lambda} = \hat{a}^\dagger_{\mathbf{k},\lambda} \hat{a}_{\mathbf{k},\lambda} thus weights each mode's contribution by its occupation number, reflecting the additive nature of photon momenta. A single-photon state |1_{\mathbf{k},\lambda}\rangle is an eigenstate of \hat{\mathbf{P}} with eigenvalue \mathbf{p} = \hbar \mathbf{k}, so the magnitude is |\mathbf{p}| = \hbar k = E/c, where E = [\hbar \omega_k](/page/H-bar) is the and \omega_k = c k relates to wave number in . This linear relation underscores the massless, relativistic character of photons. In free space, the field commutes with \hat{\mathbf{P}}, conserving the total during without sources or boundaries. This conservation holds for processes like photon or , where the net remains unchanged unless external interactions occur. The borne by field modes manifests in , where photons transfer linear to absorbing or reflecting matter, exerting a force F = \Delta p / \Delta t proportional to the rate of momentum change; for perfect absorption, the pressure is I/c with intensity I. This effect enables applications such as optical trapping and . For transverse electromagnetic modes, the aligns with the direction specified by \mathbf{k}, as the divergence-free and transversality conditions ensure no longitudinal component; this directional ties the photon's motion to its in the quantized field.

Spin and Polarization

Photons, as of the , are massless spin-1 bosons, possessing intrinsic characterized primarily by their spin degree of freedom in the free field approximation. The total operator for the field decomposes into orbital and spin contributions, \mathbf{J} = \mathbf{L} + \mathbf{S}, where for plane-wave modes propagating along a well-defined , the orbital part \mathbf{L} vanishes, leaving the spin \mathbf{S} as the dominant component aligned with the axis. In the quantized description, the spin angular momentum for a specific wave vector mode \mathbf{k} is represented by the operator \hat{\mathbf{S}}_{\mathbf{k}} = \sum_{\lambda,\lambda'} \hbar \boldsymbol{\sigma}_{\lambda\lambda'} \hat{a}^\dagger_{\mathbf{k},\lambda} \hat{a}_{\mathbf{k},\lambda'}, where \hat{a}^\dagger_{\mathbf{k},\lambda} and \hat{a}_{\mathbf{k},\lambda} are the creation and annihilation operators for photons in mode \mathbf{k} with polarization index \lambda = 1,2, and \boldsymbol{\sigma} denotes the vector of Pauli matrices acting in the two-dimensional polarization space. This formulation treats the two transverse polarization states as a pseudospin system, capturing the vectorial nature of the photon's spin. The eigenvalues of the spin projection along the propagation direction \mathbf{k}, known as , are \pm \hbar for single-photon states, corresponding to the two possible transverse polarizations. These helicity states align with s: the +\hbar eigenvalue corresponds to right-handed (where the electric field vector rotates in the right-hand sense along \mathbf{k}), and -\hbar to left-handed , reflecting the intrinsic handedness of the photon's . Due to the imposed by gauge invariance in the free , photons exhibit no scalar (spin-0) component nor longitudinal (helicity-0) modes, restricting the to the two helicity states. This exclusion arises from the Lorentz gauge and the physical , which eliminate unphysical timelike and longitudinal , ensuring only transverse modes contribute to observable states.

Rest Mass and Effective Mass

In the quantization of the electromagnetic field, photons are massless particles in , as dictated by the relativistic energy-momentum E^2 = p^2 c^2 + m^2 c^4, where E is the , p is the , c is the , and m is the rest mass. For photons, in yield the linear relation E = p c, which requires m = 0 to satisfy the for all frequencies. This zero rest mass is a fundamental consequence of the gauge invariance of the and ensures that photons propagate at the invariant speed c in , distinguishing them from massive particles. Experimental efforts to detect any nonzero photon rest mass have yielded extraordinarily tight upper bounds, primarily from astrophysical observations. Analyses of magnetohydrodynamics and galactic magnetic fields constrain the photon rest mass to m_\gamma < 10^{-18} eV/c^2 at 95% confidence level, with no evidence for a nonzero value. More recent constraints from fast radio bursts and pulsar timing in 2023–2024 maintain or slightly tighten this limit, confirming consistency with m_\gamma = 0 across diverse scales from laboratory tests to cosmic phenomena. These bounds underscore the photon's role as a truly massless excitation in the quantized field. Although the intrinsic rest mass of photons is zero, concepts of effective mass arise in certain environments, such as dielectric media or confined geometries, where the propagation is modified without altering the vacuum dispersion. In a linear dielectric medium with refractive index n > 1, the phase velocity v_p = c/n leads to an effective mass m^* = \hbar \omega / v_p^2 = n^2 \hbar \omega / c^2, reflecting the reduced speed and increased relative to for a given energy \hbar \omega. This effective describes the photon's behavior, arising from interactions with the medium's polarizable constituents, but it does not imply a true rest mass; the photon remains relativistic with zero . In optical cavities, boundary conditions quantize the modes, introducing a cutoff wavevector that curves the near zero in-plane , thereby endowing photons with an effective proportional to the inverse cavity size, enhancing light-matter coupling in . The distinction is crucial: the true rest mass remains zero, preserving the vacuum speed c, while effective masses are context-dependent and vanish in the absence of the medium or confinement.

Classical Limit

Semiclassical Approximation

The semiclassical approximation in the quantization of the electromagnetic field arises primarily through the use of coherent states, which minimize quantum fluctuations and allow the expectation values of field operators to closely mimic classical fields. In these states, the expectation value of the operator \langle \hat{\mathbf{A}} \rangle approximates the classical \mathbf{A}_{cl}(t), such that the derived electric and magnetic fields satisfy the classical equations in the presence of sources. This approximation holds because coherent states are eigenstates of the annihilation operator, leading to displaced vacuum states where the field behaves like a classical wave with added . The application of Ehrenfest's theorem further justifies this reduction to classical dynamics by governing the time evolution of expectation values for the field operators. Specifically, the theorem yields equations such as \frac{d}{dt} \langle \hat{\mathbf{E}} \rangle = c \langle \nabla \times \hat{\mathbf{B}} \rangle (in Gaussian units) and \frac{d}{dt} \langle \hat{\mathbf{B}} \rangle = -c \langle \nabla \times \hat{\mathbf{E}} \rangle , mirroring the source-free Maxwell equations when averaged over coherent states. These relations demonstrate how quantum field commutators translate into the classical curl identities for the mean fields, provided the state supports non-vanishing expectation values. This approximation is valid in the regime of large mean photon number \langle N \rangle \gg 1, where the relative quantum fluctuations become negligible. The energy uncertainty \Delta E \sim \sqrt{\langle N \rangle} \hbar \omega scales as the square root of the intensity, making the fractional fluctuation \Delta E / E \approx 1 / \sqrt{\langle N \rangle} small compared to the classical signal. However, it breaks down for low-intensity scenarios, such as single-photon states where \langle N \rangle \approx 1 and quantum effects dominate, or vacuum states exhibiting phenomena like the Casimir force due to zero-point fluctuations.

Correspondence to Classical Fields

In the classical regime of the quantized , the two-point functions of the operators approach those of a classical . Specifically, the expectation value \langle \hat{\mathbf{E}}(\mathbf{r},t) \hat{\mathbf{E}}(\mathbf{r}',t') \rangle approximates \mathbf{E}_{cl}(\mathbf{r},t) \mathbf{E}_{cl}^*(\mathbf{r}',t'), where \mathbf{E}_{cl} represents the classical , reflecting the factorization typical of coherent or high-intensity states where quantum fluctuations become negligible relative to the mean . This correspondence arises in through the analysis of normally ordered functions, as developed by Glauber, where coherent states minimize uncertainty and mimic deterministic classical waves. In the high-temperature or high-occupation-number limit, the quantum statistical description of the transitions to classical via the Bose-Einstein . When the k_B T greatly exceeds the \hbar \omega (i.e., h\nu / kT \ll 1), the occupation number \langle n \rangle = 1 / (e^{\hbar \omega / k_B T} - 1) approximates k_B T / \hbar \omega, recovering the classical Rayleigh-Jeans for the u(\nu) = (8\pi \nu^2 / c^3) k_B T. This limit demonstrates how the quantized field's aligns with classical equipartition of energy among modes, resolving the by ensuring the quantum description asymptotes to the classical Rayleigh-Jeans regime for long wavelengths or elevated temperatures. Decoherence plays a crucial role in suppressing quantum superpositions for macroscopic electromagnetic fields, facilitating the emergence of classical-like behavior. Interactions with environmental modes, such as scattering off atoms or absorption in media, rapidly entangle the field with the surroundings, leading to the exponential decay of off-diagonal elements and the effective selection of pointer states that resemble classical field configurations. For instance, in experiments with microwave fields, decoherence times on the order of microseconds or less prevent coherent superpositions of macroscopic field amplitudes, enforcing a classical appearance even though the underlying dynamics remain quantum mechanical. Historically, the correspondence between quantum radiation and classical fields was advanced by in the 1920s through his application of to electromagnetic fluctuations. In the 1926 Dreimännerarbeit with and Heisenberg, Jordan quantized the field modes as harmonic oscillators, deriving fluctuation formulas that unified wave and particle aspects and recovered classical limits via contributions that vanish in the high-energy regime. This work built on Einstein's 1909 fluctuation analysis, establishing an early for radiation that emphasized the quantum theory's recovery of classical predictions under appropriate limits.

References

  1. [1]
    [PDF] Quantization of the electromagnetic field - MIT OpenCourseWare
    Field in cavity. • Field in a cavity of volume V = LxLyLz. • Given the boundary conditions, the normal modes are: un,α = Aα cos(kn,xrx) sin(kn,yry) sin(kn ...
  2. [2]
    Quantization of the Electromagnetic Field [lam3]
    Quantization of the Electromagnetic Field [lam3]. The purpose of this elementary introduction is tailored toward the needs of quantum optics. QED will later ...
  3. [3]
    A.23 Quantization of radiation - Florida State University
    And to learn something about the effects of quantization of the electromagnetic field. Consider first a simple wave function where there are exactly $i$ photons ...
  4. [4]
    Quantum Field Theory > The History of QFT (Stanford Encyclopedia ...
    Its first achievement, namely the quantization of the electromagnetic field is “still the paradigmatic example of a successful quantum field theory” (Weinberg ...
  5. [5]
    PAM Dirac and the discovery of quantum mechanics - AIP Publishing
    Mar 1, 2011 · In short, Born and Jordan and Dirac independently discovered canonical quantization, and thereby transformed Heisenberg's scheme into a complete ...
  6. [6]
    [PDF] Quantization: History and problems - OSTI.GOV
    Feb 20, 2022 · I discuss the early history of this notion of quantization with emphasis on the works of Schrödinger and Dirac, and how quantization fit into ...
  7. [7]
    Electromagnetic field quantization and quantum optical input-output ...
    Dec 27, 2019 · Quantization of the electromagnetic field in dielectrics. Phys. Rev. A. 1992;46:4306–4321. doi: 10.1103/PhysRevA.46.4306. [DOI] [PubMed] ...
  8. [8]
    The Long Road to Maxwell's Equations - IEEE Spectrum
    Dec 1, 2014 · Maxwell may have selected that name for the field—today known as magnetic vector potential—because its derivative with respect to time yields ...
  9. [9]
    [PDF] The conceptual origins of and gauge theory - Maxwell's equations
    Nov 12, 2014 · Maxwell had chosen well: Earlier in. 1851 Thomson had introduced what we now call the vector potential A to express the magnetic field H.
  10. [10]
    Electrodynamics in Relativistic Notation - Feynman Lectures - Caltech
    The potentials at P can be computed in either frame. We will consider one example of the usefulness of the idea of the four-potential. What are the vector ...
  11. [11]
    The potential 4-vector - Richard Fitzpatrick
    The field equations which govern classical electromagnetism can all be summed up in a single 4-vector equation.
  12. [12]
    [PDF] Theory of electromagnetic fields
    6.1 Relationships between the potentials and the fields. The scalar potential φ and vector potential ~A are defined so that the electric and magnetic fields are.
  13. [13]
  14. [14]
    [PDF] Electromagnetic Potentials and Gauge Invariance Maxwell's ...
    Apr 1, 2015 · There are two common choices, the Lorenz gauge, and the. Coulomb gauge. Lorenz Gauge. For the Lorenz Gauge we impose the gauge constraint: 1 c.
  15. [15]
    [PDF] arXiv:2003.07473v4 [hep-th] 22 Nov 2021
    Nov 22, 2021 · Indeed Coulomb gauge quantization goes together with two possible views about the ontological status of the longitudinal modes of the electric ...
  16. [16]
    [PDF] Quantization of the Free Electromagnetic Field: Photons and Operators
    The electric field is proportional to the canonical momentum, E = −4πcp. So really, the electric field energy term already looks like a sum of squared momenta.
  17. [17]
    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 ...
  18. [18]
    [PDF] Quantum Field Theory - DAMTP
    We will show how photons arise from the quantization of the electromagnetic field and how massive, charged particles such as ... on canonical quantization.<|control11|><|separator|>
  19. [19]
    [PDF] 10. The electromagnetic field - MIT OpenCourseWare
    The solution of the wave equation can thus be facilitated by representing the electric field as a sum of normal mode functions: E(x, t) = fm(t)um(x). mfm(t)=0. ...
  20. [20]
    The Quantum Theory Of Radiation Ed. 2 : Heitler, W. - Internet Archive
    Jan 20, 2017 · The Quantum Theory Of Radiation Ed. 2. by: Heitler, W ... PDF WITH TEXT download · download 1 file · SINGLE PAGE PROCESSED JP2 ZIP ...
  21. [21]
    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.
  22. [22]
    [PDF] Planck's Law and Light Quantum Hypothesis.
    Planck's Law and Light Quantum Hypothesis. S.N. Bose. (Received 1924). Planck's formula for the distribution of energy in the radiation from a black body ...
  23. [23]
    [PDF] Quantization of electromagnetic field Masatsugu Sei Suzuki ...
    Feb 22, 2017 · Momentum operator for photon. The momentum density (erg s/cm4) of an electromagnetic field is the Poynting vector. ),. (. ˆ. ),. (. ˆ. 4. ),. (.Missing: ħk | Show results with:ħk
  24. [24]
    Single-photon quantum regime of artificial radiation pressure on a ...
    Mar 17, 2020 · Electromagnetic fields carry momentum, which upon reflection on matter gives rise to the radiation pressure of photons.
  25. [25]
    [PDF] 6. Quantum Electrodynamics - DAMTP
    Coulomb gauge is sometimes called radiation gauge. 3Named after Lorenz who had the misfortune to be one letter away from greatness.
  26. [26]
    [2004.03771] Quantum field theory for spin operator of the photon
    Apr 8, 2020 · Our work shows that the spin and OAM operators commute which is important for simultaneously observing and separating the SAM and OAM.
  27. [27]
    [1011.3608] Photon spin operator and Pauli matrix - arXiv
    Nov 16, 2010 · The spin operator \hat{\boldsymbol \gamma} defined on the space of unit spinors, referred to as the Jones space, has only component along the wave vector.Missing: electrodynamics | Show results with:electrodynamics
  28. [28]
    [PDF] γ (photon) - Particle Data Group
    May 31, 2024 · 2WANG 23B use fast radio burst photon mass dependent dispersion relation to determine an upper limit of the photon mass. 3BONETTI 17 uses ...Missing: rest | Show results with:rest
  29. [29]
    [1801.00679] An Effective Photon Momentum in a Dielectric Medium
    Nov 24, 2017 · We use a relativistic argument to define an effective photon that travels through a transparent (non-absorbing) nondispersive dielectric medium ...Missing: seminal mass
  30. [30]
    Effective mass in cavity QED | Phys. Rev. A
    Based on the band theory, we introduce a set of effective mass parameters that approximately describe the effect of the cavity on the atomic motion, with the ...Article Text · INTRODUCTION · PROPAGATION OF WAVE... · CONCLUSION
  31. [31]
    [PDF] Quantum Field Theory IV (Radiation Field) - 221B Lecture Notes
    ... coherent state, which has an expectation value for the vector potential. This expectation value is nothing but what we normally obtain by solving Maxwell's ...
  32. [32]
    Classical and Quantum Mechanical Correlation Functions of Fields ...
    Aug 8, 1970 · The two-point, two-time correlation functions of classical and quantum mechanical fields in thermal equilibrium in an arbitrary domain are ...Missing: correspondence | Show results with:correspondence
  33. [33]
    [PDF] Correlation functions in optics and quantum optics - UMD Physics
    Correlation functions tell us something about fluctuations. The correlation functions have classical limits. They are related to conditional measurements. They ...Missing: correspondence | Show results with:correspondence
  34. [34]
    [PDF] Roy J. Glauber - Nobel Lecture
    So, according to Dirac, the electromagnetic field is made up of field am- plitudes that can oscillate harmonically. But these amplitudes, because of the ever- ...
  35. [35]
    [PDF] Einstein's Fluctuation Formula. A Historical Overview - arXiv
    This way we would end up with the Rayleigh-Jeans law, u. R-J=(8πν2. /c. 3 ... the Rayleigh-Jeans formula (in the limit hν/kT<<1) from the photon concept ...
  36. [36]
    [PDF] Decoherence and the Transition from Quantum to Classical ... - arXiv
    The application of quantum physics to information processing has also transformed the nature of interest in the process of decoherence: At the time of my orig-.
  37. [37]
    [PDF] Pascual Jordan's resolution of the conundrum of the wave-particle ...
    Sep 24, 2007 · In 1909, Einstein derived a formula for the mean square energy fluctuation in a subvolume of a box filled with black-body radiation.