Fact-checked by Grok 2 weeks ago

Photon polarization

Photon polarization is a fundamental quantum mechanical property of photons, the of , describing the orientation of the associated perpendicular to the direction of propagation. In classical , polarization arises from the coherent of the in a specific direction, but at the quantum level, it manifests as the intrinsic or of the , which can take values of ±ħ along the propagation axis. This property enables photons to exist in various states, including linear polarizations (such as or vertical, defined relative to a chosen basis) and circular polarizations (right-handed or left-handed, corresponding to clockwise or counterclockwise rotation of the field ). Unlike classical waves, where is deterministic, photon polarization exhibits probabilistic behavior upon measurement, as described by and the projection postulate. For instance, a in a linearly polarized at an angle θ to a 's axis has a transmission probability of cos²θ through that , illustrating Malus's law in a quantum context. , in quantum terms, consists of an incoherent mixture of orthogonal polarization states, resulting in a 50% average transmission probability through any linear . Polarization states form a two-dimensional for a given propagation direction, spanned by basis states like |H⟩ (horizontal) and |V⟩ (vertical), allowing general states to be expressed as superpositions such as |ψ⟩ = cosθ |H⟩ + sinθ |V⟩ for at angle θ. Circular states, |R⟩ and |L⟩, are eigenstates of and can be written as superpositions of linear states: |R⟩ = \frac{1}{\sqrt{2}} (|H⟩ + i |V⟩) and |L⟩ = \frac{1}{\sqrt{2}} (|H⟩ - i |V⟩). of collapses the superposition into one of the basis states, highlighting the role of observation in . In , photon polarization underpins phenomena such as entanglement, where paired photons exhibit correlated polarizations that violate classical inequalities, enabling applications in processing. It is also crucial for protocols like , which use states to encode qubits, and for experiments demonstrating superposition, such as photons passing through crossed with an intermediate 45° , where 25% transmission occurs due to quantum . Birefringent materials, like , further exploit polarization by splitting light into and rays based on the angle relative to the optic axis, with refractive indices n_o ≈ 1.658 and n_e ≈ 1.486.

Classical Polarization of Electromagnetic Waves

Linear, Circular, and Elliptical Polarization States

of electromagnetic refers to the orientation and behavior of the vector as the wave propagates. In classical , are transverse, meaning their electric and oscillate perpendicular to the direction of propagation. This transverse nature restricts to the plane normal to the wave's path, allowing for distinct states based on the field's pattern. Linear polarization occurs when the electric field vector oscillates along a fixed straight line within the transverse plane. For instance, horizontal linear polarization has the field vibrating parallel to the ground, while vertical linear polarization aligns it perpendicular to the ground. This state arises from a single component of the field dominating or when orthogonal components are in phase. Circular polarization describes a situation where the vector rotates at a constant magnitude in the transverse plane as the wave advances, tracing a . Right-handed (or right-circular) polarization involves clockwise rotation when viewed facing the oncoming wave, whereas left-handed (or left-circular) polarization rotates counterclockwise. The field completes one full rotation per along the propagation direction. Elliptical polarization represents the most general form, where the electric field traces an ellipse in the transverse plane, combining elements of linear and circular motion. The ellipse is characterized by its major and minor axes, which indicate the varying amplitudes of the orthogonal field components, and a tilt angle defining the ellipse's orientation relative to reference axes. This state results from the superposition of two orthogonal linear components with a phase difference between 0° and 180° that is neither 0°/180° (linear) nor exactly 90° with equal amplitudes (circular). Linear and circular polarizations are special cases of elliptical polarization, with the ellipse degenerating to a line or circle, respectively. The concept of as evidence for the nature of was established in the 1820s by , who through experiments on and demonstrated that light's vibrations occur in planes perpendicular to its propagation direction. Fresnel's work, including the 1822 discovery of via decomposition of linearly polarized light, solidified the transverse model over longitudinal alternatives.

Mathematical Representation Using Jones Vectors

The polarization state of a fully coherent, monochromatic plane electromagnetic wave propagating along the z-axis can be described using Jones vectors, a introduced by R. C. Jones in 1941. The Jones vector is a two-component complex column representing the amplitudes in the horizontal (H) and vertical (V) basis: \mathbf{E} = \begin{pmatrix} E_H \\ E_V \end{pmatrix}, where E_H and E_V are complex numbers encoding both the magnitudes and relative phase of the field components. For fully polarized light, the Jones vector is typically normalized such that |E_H|^2 + |E_V|^2 = 1, ensuring the vector lies on the unit circle in the ; the relative phase difference \delta = \arg(E_V) - \arg(E_H) then determines whether the is linear (\delta = 0 or \pi), circular (\delta = \pm \pi/2), or elliptical (general \delta). Specific polarization states correspond to standard normalized Jones vectors in the H-V basis. Horizontal linear polarization is represented by \begin{pmatrix} 1 \\ 0 \end{pmatrix}, vertical linear by \begin{pmatrix} 0 \\ 1 \end{pmatrix}, and 45° linear (diagonal) by \frac{1}{\sqrt{2}} \begin{pmatrix} 1 \\ 1 \end{pmatrix}. Right-circular polarization, defined for light propagating toward the observer with the electric field rotating clockwise, is given by \frac{1}{\sqrt{2}} \begin{pmatrix} 1 \\ -i \end{pmatrix}, while left-circular is \frac{1}{\sqrt{2}} \begin{pmatrix} 1 \\ i \end{pmatrix}. Conversions between states, such as from linear to circular polarization, can be achieved by introducing a relative phase shift of \pi/2; for example, applying a quarter-wave retarder to horizontal linear light (\begin{pmatrix} 1 \\ 0 \end{pmatrix}) with fast axis at 45° yields the right-circular state. Under a rotation of the coordinate system by an angle \theta (counterclockwise when looking toward the source), the Jones vector transforms via the unitary : R(\theta) = \begin{pmatrix} \cos \theta & \sin \theta \\ -\sin \theta & \cos \theta \end{pmatrix}, such that the new vector is R(\theta) \mathbf{E}; this preserves the polarization ellipse's shape and orientation relative to the new axes. For instance, rotating horizontal linear polarization by 45° yields the diagonal state: R(45^\circ) \begin{pmatrix} 1 \\ 0 \end{pmatrix} = \frac{1}{\sqrt{2}} \begin{pmatrix} 1 \\ 1 \end{pmatrix}. The Jones vector formalism is limited to fully polarized, coherent plane waves and does not apply to partially polarized or incoherent light, for which the Mueller-Stokes calculus is required instead.

Geometric Visualization of Polarization

The Poincaré sphere provides a geometric representation of the polarization states of fully polarized light on the surface of a unit sphere in three-dimensional space. Developed by Henri Poincaré in the late 19th century as a tool for analyzing the evolution of light's polarization through interactions with matter, the sphere maps elliptical polarization states using spherical coordinates related to the orientation and ellipticity of the polarization ellipse. The north pole corresponds to right-circular polarization, the south pole to left-circular polarization, the equator to all linear polarization states (with horizontal and vertical polarizations at opposite points), and latitudes away from the equator represent elliptical states, where the angle from the equator quantifies the degree of ellipticity. Jones vectors, which algebraically describe polarization in the linear basis, map to points on the via from the complex plane of normalized Jones vectors to the sphere's surface. This projection identifies the Jones vector components with coordinates on the sphere, and the Cartesian axes (S₁, S₂, S₃) of the sphere directly correspond to the normalized , satisfying S₁² + S₂² + S₃² = 1 for fully polarized light: \begin{align*} S_1 &= \cos(2\chi) \cos(2\psi), \\ S_2 &= \cos(2\chi) \sin(2\psi), \\ S_3 &= \sin(2\chi), \end{align*} where ψ is the orientation angle and χ is the ellipticity angle of the polarization ellipse. Polarization transformations, such as those induced by wave plates (retarders) or optical rotators, appear as rotations of points on the sphere, with great circles tracing the paths of evolving states under linear retardance or rotation. For instance, a quarter-wave retarder rotates states around the S₁ axis by 90°, converting linear to circular polarization. This geometric view simplifies the analysis of sequential optical elements, as compositions of transformations correspond to successive rotations on the sphere. The Poincaré sphere's advantages lie in its intuitive visualization of complex elliptical states and its extension to partially polarized light, where points inside the sphere (with S₁² + S₂² + S₃² < 1) represent mixtures of polarization states. It also serves as a natural introduction to the , which generalizes to describe transformations for both polarized and unpolarized light using 4×4 matrices that act as rotations and scalings in this space, though without delving into matrix details here.

Physical Quantities in Polarized Electromagnetic Waves

Energy Density and Distribution

The energy density of an electromagnetic wave in free space is given by the sum of the electric and magnetic field contributions, u = \frac{1}{2} \epsilon_0 | \mathbf{E} |^2 + \frac{1}{2 \mu_0} | \mathbf{B} |^2, where \epsilon_0 and \mu_0 are the permittivity and permeability of free space, respectively. For plane waves propagating in vacuum, the magnitudes of the electric and magnetic fields are related by | \mathbf{B} | = | \mathbf{E} | / c, where c = 1 / \sqrt{\epsilon_0 \mu_0} is the speed of light, leading to equal partitioning of the energy between the electric and magnetic fields: u = \epsilon_0 | \mathbf{E} |^2 = | \mathbf{B} |^2 / \mu_0. This equal partition holds regardless of the wave's polarization state, as the total energy density depends on the overall field strengths. Polarization influences the distribution of energy within the electric field components transverse to the propagation direction. For a plane wave propagating along the z-axis, the electric field can be decomposed into x- and y-components, \mathbf{E} = (E_x, E_y, 0), such that | \mathbf{E} |^2 = | E_x |^2 + | E_y |^2. The fraction of energy associated with each component is proportional to the square of its amplitude; for instance, in a linearly polarized wave at 45° to the x-axis, | E_x | = | E_y | = | \mathbf{E} | / \sqrt{2}, resulting in 50% of the electric energy in each component. In circular polarization, the amplitudes are equal but phases differ by 90°, maintaining the same total energy density while distributing it equally over time between the components. The flow of this energy is described by the Poynting vector, \mathbf{S} = \mathbf{E} \times \mathbf{H}, which represents the instantaneous power flux density. For monochromatic plane waves, the time-averaged intensity is \langle \mathbf{S} \rangle = \frac{1}{2} \Re ( \mathbf{E} \times \mathbf{H}^* ), directed along the propagation axis. In free space, \mathbf{H} = \mathbf{E} / Z_0, where Z_0 = \sqrt{\mu_0 / \epsilon_0} \approx 377 \, \Omega is the impedance of free space, yielding \langle S \rangle = \frac{1}{2} \epsilon_0 c | \mathbf{E} |^2. Notably, for a fixed total electric field magnitude | \mathbf{E} |, the time-averaged energy flux is independent of polarization, as it depends only on the overall intensity rather than the orientation or ellipticity of the field components.

Momentum Density

In classical electromagnetism, the linear momentum density \mathbf{g} of electromagnetic fields is given by \mathbf{g} = \frac{1}{c^2} \mathbf{S}, where \mathbf{S} is the representing the energy flux and c is the speed of light in vacuum. This expression arises from the relativistic for the electromagnetic field, linking momentum directly to the flow of energy. For time-harmonic fields, the time-averaged momentum density follows similarly from the real part of the complex . For a plane electromagnetic wave propagating in vacuum, the magnitude of the momentum density simplifies to g = \frac{u}{c}, where u is the energy density of the wave. This result holds because the Poynting vector magnitude S = c u for such waves, leading to g = \frac{S}{c^2} = \frac{u}{c} directed along the propagation axis. Notably, g is independent of the polarization state—whether linear, circular, or elliptical—as long as the wave intensity (and thus u) remains the same, since the transverse orientation of the electric and magnetic fields does not affect the longitudinal energy flow in isotropic media. The momentum carried by polarized waves manifests physically through radiation pressure, the force exerted upon momentum transfer to matter. For a plane wave at normal incidence on a perfectly absorbing surface, the time-averaged pressure is P = \frac{I}{c}, where I is the intensity (equal to the time-averaged S). This pressure arises from the complete absorption of the wave's momentum flux, with no dependence on polarization in isotropic absorbers. In structured or anisotropic media, however, polarization can influence momentum transfer via spin-momentum locking, where the wave's spin (related to ) couples to its linear momentum direction, leading to transverse shifts or directional selectivity in the classical limit. Conservation of total linear momentum in electromagnetic systems requires considering both field and mechanical contributions, as isolated field momentum is not conserved. For a localized electromagnetic wave packet, the total field momentum is the volume integral \mathbf{P} = \int \mathbf{g} \, dV = \frac{1}{c^2} \int \mathbf{S} \, dV, which balances any mechanical momentum changes during interactions like emission or absorption. This integral form ensures overall momentum conservation in closed systems, consistent with applied to space-translation symmetry in .

Angular Momentum Density

The spin angular momentum density of electromagnetic fields originates from the rotational character of the electric and magnetic field vectors, particularly in non-linear polarization states. For circularly polarized plane waves, this density manifests along the direction of propagation due to the helical rotation of the field vectors as the wave advances, imparting a torque on absorbing matter. A general expression for the spin angular momentum density is \mathbf{l}_s = \epsilon_0 \Im (\mathbf{E}^* \times \mathbf{A}), where \epsilon_0 is the vacuum permittivity, \mathbf{E} is the complex electric field, and \mathbf{A} is the vector potential; however, for plane waves, it simplifies to l_z = \pm \frac{u}{\omega}, with u denoting the energy density and \omega the angular frequency, the sign indicating right- or left-handed circular polarization. In linearly polarized waves, the spin angular momentum density is zero, as the field vectors oscillate without net rotation. For circular polarization, it reaches its maximum value of \pm u / \omega, while elliptical polarization yields an intermediate magnitude proportional to the degree of circularity \sigma (where |\sigma| \leq 1), given by l_z = \sigma \frac{u}{\omega}. The total angular momentum density of the field includes both spin and orbital components, but for uniformly polarized plane waves, the orbital contribution vanishes, leaving the spin as the sole contributor. Experimental verification of spin angular momentum transfer has been achieved using optical tweezers, where circularly polarized light rotates trapped birefringent particles, such as calcite microspheres, at rates consistent with the absorption of spin angular momentum from the beam. In these setups, the rotation direction and speed align with the polarization handedness, confirming the classical torque exerted by the field's spin density on matter.

Interactions of Polarized Waves with Matter

Passage Through Polarizing Filters

When a beam of polarized electromagnetic radiation encounters an ideal linear polarizing filter, the filter transmits the component of the electric field vector parallel to its transmission axis while absorbing the orthogonal component. This interaction is governed by Malus's law, which states that the transmitted intensity I is related to the incident intensity I_0 by I = I_0 \cos^2 \theta, where \theta is the angle between the polarization direction of the incident light and the filter's axis. Étienne-Louis Malus formulated this law in 1810 based on experiments with reflected light and calcite crystals, establishing the foundational principle for intensity variation in polarized light transmission. In the Jones calculus framework, the action of an ideal linear polarizer can be represented by a projection matrix that selects the parallel field component. For a horizontal polarizer (transmission axis along the x-direction), the Jones matrix is \begin{pmatrix} 1 & 0 \\ 0 & 0 \end{pmatrix}, which, when multiplied by the input Jones vector describing the incident field's horizontal (E_x) and vertical (E_y) components, yields the transmitted field as \begin{pmatrix} E_x \\ 0 \end{pmatrix}. For a polarizer at an arbitrary angle \phi, the matrix is obtained by rotating the horizontal form using the rotation matrix R(\phi) = \begin{pmatrix} \cos \phi & \sin \phi \\ -\sin \phi & \cos \phi \end{pmatrix}, resulting in J_p(\phi) = R(\phi) \begin{pmatrix} 1 & 0 \\ 0 & 0 \end{pmatrix} R(-\phi). This formalism, introduced by R. Clark Jones in 1941, facilitates the modeling of sequential optical elements by matrix multiplication. The absorbed portion of the incident energy, corresponding to I_0 \sin^2 \theta, is converted into heat within the filter material, ensuring conservation of energy as thermal dissipation. For unpolarized incident light, which has equal intensities in all polarization directions, an ideal linear polarizer transmits on average 50% of the incident intensity, as the random orientations average to \langle \cos^2 \theta \rangle = 1/2. In contrast, for linearly polarized input aligned with the axis (\theta = 0), transmission is complete (I = I_0), while crossed alignment (\theta = 90^\circ) results in zero transmission. Real-world polarizing filters, such as Polaroid sheets, approximate ideal behavior through dichroic absorption. These consist of polyvinyl alcohol films stretched to align polymer chains, into which iodine molecules are embedded, preferentially absorbing light polarized perpendicular to the alignment direction due to resonance with the molecular transitions. Edwin H. Land developed this H-type polarizer in 1938, enabling practical, low-cost polarization control for applications like sunglasses and optical instruments. While ideal models assume perfect transmission and absorption, actual devices exhibit slight imperfections, such as minor leakage (extinction ratios typically around 10^2 to 10^3) and wavelength-dependent efficiency, but they faithfully reproduce over visible and near-infrared ranges.

Propagation in Birefringent Crystals

Birefringent crystals exhibit an optical property known as birefringence, where the refractive index depends on the polarization direction of the incident light, leading to double refraction. In uniaxial crystals such as , light propagating along a direction not aligned with the optic axis splits into two orthogonally polarized rays: the ordinary ray (o-ray), which experiences a constant refractive index n_o regardless of direction, and the extraordinary ray (e-ray), which has a direction-dependent refractive index n_e that differs from n_o. For example, in , n_o = 1.658 and n_e = 1.486 at 589 nm, resulting in a birefringence magnitude |n_e - n_o| \approx 0.172. When linearly polarized light enters a birefringent crystal at an angle to the optic axis, it decomposes into orthogonal o- and e-ray components that propagate with different velocities due to their distinct refractive indices, a phenomenon called double refraction. The o-ray follows the standard laws of refraction, while the e-ray deviates slightly, causing spatial separation of the beams inside the crystal. This splitting occurs because the crystal's anisotropic structure imposes different phase velocities on the two polarization components, without absorption. As the o- and e-rays traverse the crystal of thickness d, they accumulate a relative phase difference \delta = \frac{2\pi d}{\lambda} (n_o - n_e), where \lambda is the wavelength in vacuum. This phase retardation alters the output polarization state; for instance, an input linear polarization can emerge as elliptically polarized light, with the ellipticity depending on \delta and the input orientation relative to the optic axis. The propagation through a birefringent retarder can be described using Jones calculus, where the Jones matrix for a linear retarder aligned with the principal axes is diagonal: \begin{pmatrix} e^{i\delta/2} & 0 \\ 0 & e^{-i\delta/2} \end{pmatrix}, assuming equal transmission amplitudes and a common phase factor. This matrix applies a differential phase shift between the o- and e-components. A specific case is the quarter-wave plate, where \delta = \pi/2, which converts linearly polarized light incident at 45° to the fast axis into circularly polarized light by equalizing the amplitudes while introducing a 90° phase difference. Waveplates, fabricated from birefringent crystals like quartz or mica, exploit this effect for precise polarization control in optical systems, such as rotating linear polarization or generating specific elliptical states. Historically, the Wollaston prism, invented by William Hyde Wollaston around 1803, uses two cemented calcite prisms to spatially separate o- and e-rays, enabling applications in polarimetry and beam splitting.

Energy Conservation in Optical Interactions

In the Jones calculus framework, energy conservation in lossless optical interactions involving polarized light is ensured through unitary transformations represented by Jones matrices. A unitary Jones matrix U satisfies U^\dagger U = I, where U^\dagger is the conjugate transpose and I is the identity matrix, preserving the norm of the input Jones vector \mathbf{E}_{in}. For an output vector \mathbf{E}_{out} = U \mathbf{E}_{in}, the squared magnitude |\mathbf{E}_{out}|^2 = \mathbf{E}_{out}^\dagger \mathbf{E}_{out} = \mathbf{E}_{in}^\dagger U^\dagger U \mathbf{E}_{in} = |\mathbf{E}_{in}|^2, which corresponds to the total intensity or energy flux of the electromagnetic wave remaining unchanged. This norm preservation reflects the absence of absorption or dissipation in the system, applicable to devices such as waveplates or retarders that alter polarization without energy loss. Hermitian operators play a complementary role in describing observables within these interactions; for instance, the intensity measurement operator is self-adjoint (H = H^\dagger), guaranteeing real eigenvalues that quantify measurable energy distributions, such as the projected intensity along a specific polarization axis. A practical example is the interaction of linearly polarized light with a birefringent crystal acting as a unitary retarder. An input Jones vector representing horizontal linear polarization, \mathbf{E}_{in} = \begin{pmatrix} 1 \\ 0 \end{pmatrix}, transforms via the retarder's Jones matrix—typically diagonal with phase shifts e^{i\delta/2} and e^{-i\delta/2} along the fast and slow axes—into an elliptical polarization state, such as circular for a quarter-wave retardance (\delta = \pi/2). Despite the change in polarization geometry, the total energy remains conserved, as |\mathbf{E}_{out}|^2 = 1, matching the input. Unitary transformations also maintain the orthogonality of dual polarization states, where "dual" refers to pairs of orthogonal basis vectors, such as horizontal and vertical linear polarizations. If two input states \mathbf{E}_1 and \mathbf{E}_2 satisfy \mathbf{E}_1^\dagger \mathbf{E}_2 = 0, their outputs U \mathbf{E}_1 and U \mathbf{E}_2 retain this property, with swapped or rotated components but no cross-energy transfer, upholding conservation in multi-component systems. In contrast, absorbing elements like ideal polarizers introduce non-unitary Jones matrices, incorporating explicit loss terms that reduce the output norm; for a horizontal polarizer, the matrix \begin{pmatrix} 1 & 0 \\ 0 & 0 \end{pmatrix} projects the field, halving the intensity for diagonal input polarization and violating unitarity due to energy dissipation as heat.

Quantum Description of Photon Polarization

Energy, Momentum, and Spin of Single Photons

In quantum electrodynamics, the energy of a single photon is given by E = h \nu = \hbar \omega, where h is , \nu is the frequency, \hbar = h / 2\pi is the , and \omega = 2\pi \nu is the angular frequency; this energy is independent of the photon's polarization state. The quantization of light energy into discrete photon packets was first proposed by to explain the , establishing photons as fundamental quanta of the electromagnetic field. The linear momentum of a photon is \mathbf{p} = \hbar \mathbf{k}, where \mathbf{k} is the wave vector with magnitude k = 2\pi / \lambda and \lambda the wavelength, yielding a momentum magnitude p = h \nu / c = E / c directed along the propagation direction; like energy, this momentum does not depend on polarization. This relation was experimentally confirmed through the , where X-ray photons scatter off electrons, transferring momentum consistent with particle-like behavior. Photons, as massless spin-1 bosons, possess intrinsic spin angular momentum with helicity—the projection of spin along the momentum direction—taking values \pm \hbar, corresponding to right- and left-handed circular polarization states, respectively. These two helicity eigenstates fully describe the transverse polarization degrees of freedom for photons in , arising from the representation theory of the . In the classical limit of many photons, the total energy and momentum densities recover the , while the spin angular momentum density aligns with the field's circular polarization components.

Probability Amplitudes and Quantum States

In the quantum mechanical description of photon polarization, the state is represented in a two-dimensional Hilbert space with orthonormal basis vectors |H⟩ and |V⟩ corresponding to horizontal and vertical linear polarizations, respectively. This basis captures the two possible helicity states of the photon's spin angular momentum along the propagation direction, as discussed in prior sections on single-photon properties. A general polarization state for a single photon is a coherent superposition given by |\psi\rangle = \alpha |H\rangle + \beta |V\rangle, where \alpha and \beta are complex coefficients satisfying the normalization condition |\alpha|^2 + |\beta|^2 = 1 to ensure the total probability is unity. Such superpositions enable the photon to exist simultaneously in multiple polarization configurations until measured, embodying the principle of . Upon measurement in the {|H⟩, |V⟩} basis—typically via a polarizing beam splitter or filter—the probability of detecting horizontal polarization is P(H) = |\alpha|^2, and vertical polarization is P(V) = |\beta|^2, as dictated by the . This probabilistic outcome arises because the measurement collapses the superposition to one of the basis states, with the squared modulus of the amplitude determining the likelihood. Single-photon experiments vividly illustrate these probability amplitudes, particularly in setups where polarization serves as which-path information. In a , mutually perpendicular polarizers placed over the slits (vertical over one and horizontal over the other) imprint orthogonal polarization states on photons traversing each path, rendering the paths distinguishable and suppressing the interference pattern in favor of classical single-slit diffraction profiles. If the incident photons are in a polarization superposition, the detection probabilities at the screen reflect the |α|^2 and |β|^2 amplitudes, modulated by the path-encoded states, demonstrating how measurement outcomes encode the quantum interference. For single photons, the focus remains on pure states, which preserve full phase coherence and allow maximal superposition effects, unlike mixed states that arise in incoherent ensembles and lack such predictable interference. Preparation techniques, such as spontaneous parametric down-conversion, routinely generate single photons in these pure polarization states for precise control in quantum optics experiments.

Unitary and Hermitian Operators in Polarization Measurements

In quantum optics, unitary operators describe the evolution of photon polarization states under lossless transformations, such as those induced by wave plates or beam splitters, ensuring the preservation of total probability and energy. These operators satisfy U^\dagger U = I, where I is the identity, guaranteeing that the norm of the state vector remains unchanged: if |\psi\rangle is the initial state, then \|\ U |\psi\rangle \ \| = 1. For example, a rotation of the polarization by an angle \theta around the propagation axis is represented by the unitary operator U(\theta) = \exp(-i \theta \sigma_y / 2), where \sigma_y = \begin{pmatrix} 0 & -i \\ i & 0 \end{pmatrix} is the Pauli-y matrix in the horizontal-vertical basis. This operator can be explicitly written as U(\theta) = \begin{pmatrix} \cos(\theta/2) & -\sin(\theta/2) \\ \sin(\theta/2) & \cos(\theta/2) \end{pmatrix}, and is physically realized by a half-wave plate oriented at angle \theta/2 relative to the horizontal. Similarly, a polarizing beam splitter acts as a unitary transformation that separates orthogonal polarizations into different spatial modes while preserving the overall photon number. Hermitian operators, which are self-adjoint (H = H^\dagger), represent observables in polarization measurements, yielding real eigenvalues that correspond to measurable outcomes such as detection probabilities. For a projective measurement in a specific polarization basis, the operator is a projector onto the desired state, such as H = |H\rangle\langle H| for horizontal polarization, where |H\rangle = \begin{pmatrix} 1 \\ 0 \end{pmatrix}. This operator has eigenvalues 1 (for the horizontal eigenstate) and 0 (for the orthogonal vertical state), reflecting the binary outcome of transmission or absorption in a polarizer. The expectation value \langle H \rangle = \langle \psi | H | \psi \rangle gives the probability of detecting a horizontally polarized photon. In the circular basis, which serves as the eigenbasis for the photon's helicity or spin angular momentum along the propagation direction, the states are |R\rangle = \frac{1}{\sqrt{2}} \begin{pmatrix} 1 \\ -i \end{pmatrix} (right-circular, eigenvalue +1 for spin) and |L\rangle = \frac{1}{\sqrt{2}} \begin{pmatrix} 1 \\ i \end{pmatrix} (left-circular, eigenvalue -1), with the corresponding projectors |R\rangle\langle R| and |L\rangle\langle L|. These bases are complete and orthogonal, allowing measurements of spin components. According to the measurement postulate of quantum mechanics, performing a measurement with a Hermitian operator H collapses the state |\psi\rangle to one of its eigenvectors |e_k\rangle with probability | \langle e_k | \psi \rangle |^2, and the observable's value is the corresponding eigenvalue. Post-measurement, the state updates to |e_k\rangle (normalized), altering the polarization description for subsequent operations. For instance, measuring in the horizontal basis with projector H on a state |\psi\rangle = \cos\alpha |H\rangle + \sin\alpha |V\rangle yields horizontal polarization with probability \cos^2\alpha, collapsing to |H\rangle. This process is irreversible, unlike unitary evolutions. In classical optics, unitary operators analogize to lossless transformations like rotations that preserve wave intensity, while Hermitian projectors correspond to real-valued intensity measurements after filters, where outcomes reflect absorbed or transmitted power without quantum superposition.

Uncertainty Principle Applications

The Heisenberg uncertainty principle, in its general form for two non-commuting observables represented by Hermitian operators \hat{A} and \hat{B}, states that the product of their standard deviations satisfies \Delta A \Delta B \geq \frac{1}{2} |\langle [\hat{A}, \hat{B}] \rangle|, where [\hat{A}, \hat{B}] = \hat{A}\hat{B} - \hat{B}\hat{A} is the commutator and the expectation value is taken over the quantum state. This relation imposes fundamental limits on the simultaneous measurability of incompatible properties, arising from the non-commutativity inherent in quantum mechanics. In the context of photon polarization, this principle applies to the quantum operators describing polarization states, preventing precise joint knowledge of certain polarization components. For photon polarization, the Stokes operators \hat{S}_1, \hat{S}_2, and \hat{S}_3—which generalize the classical Stokes parameters to the quantum domain—obey the commutation relations [\hat{S}_1, \hat{S}_2] = 2i \hat{S}_3 and cyclic permutations thereof, forming an \mathfrak{su}(2) algebra. Consequently, the uncertainty relation for, say, \hat{S}_1 (measuring the difference between horizontal and vertical linear polarizations) and \hat{S}_2 (measuring the difference between +45° and -45° linear polarizations) is \Delta S_1 \Delta S_2 \geq |\langle \hat{S}_3 \rangle|, where \hat{S}_3 corresponds to the circular polarization component. This implies that knowledge of the ellipticity (quantified by \langle \hat{S}_3 \rangle) introduces unavoidable uncertainty in the linear polarization components; for a state with significant circular polarization (large |\langle \hat{S}_3 \rangle|), one cannot precisely determine both \langle \hat{S}_1 \rangle and \langle \hat{S}_2 \rangle simultaneously. For a single photon, the maximum value of |\langle \hat{S}_3 \rangle| is 1 (in units where the total photon number is normalized), saturating the bound in pure states like right- or left-circular polarization. The spin angular momentum of a photon, associated with its polarization, provides another direct application. The components \hat{L}_x, \hat{L}_y, and \hat{L}_z (transverse and longitudinal spin) satisfy [\hat{L}_x, \hat{L}_y] = i \hbar \hat{L}_z and cyclic relations, leading to the uncertainty relation \Delta L_x \Delta L_y \geq \frac{\hbar}{2} |\langle L_z \rangle|. Here, \langle L_z \rangle = \pm \hbar for a circularly polarized photon, with zero variance in L_z, but the transverse components exhibit uncertainties \Delta L_x = \Delta L_y = \hbar / \sqrt{2}, exactly meeting the bound. This underscores the impossibility of knowing the full linear and circular polarization precisely at once, as linear polarization relates to transverse spin components while circular corresponds to the longitudinal one. These uncertainty relations manifest experimentally in single-photon interferometry, where polarization measurements impose limits on phase sensitivity and interference visibility. For instance, in a with polarized single photons, attempting to resolve which-path information via polarization projections disturbs the transverse coherence, reducing visibility in accordance with the uncertainty bounds on ; experiments have verified such trade-offs, with noise in linear polarization components limiting ellipticity discrimination to the quantum limit of \Delta S_1 \Delta S_2 \geq 1 for single-photon states.

Advanced Topics in Photon Polarization

Stokes Parameters for Partial Polarization

The Stokes parameters provide a mathematical framework for describing the polarization state of light that extends beyond fully polarized waves to include partially polarized and unpolarized light, representing the light as an incoherent superposition of polarized components. Introduced by in 1852, these parameters are particularly useful for analyzing real-world optical phenomena where coherence is incomplete, such as scattered light in natural environments. The Stokes vector is defined as \mathbf{S} = (S_0, S_1, S_2, S_3), where S_0 represents the total intensity of the light beam, S_1 quantifies the difference between horizontally and vertically linearly polarized components, S_2 measures the difference between linearly polarized components at +45° and -45°, and S_3 captures the difference between right- and left-circularly polarized components. These parameters are derived from time-averaged intensities measurable in the laboratory and form a real-valued vector that fully characterizes the polarization state for any degree of coherence. For fully polarized light, the relation S_0^2 = S_1^2 + S_2^2 + S_3^2 holds, but for partial polarization, S_0^2 > S_1^2 + S_2^2 + S_3^2. The degree of polarization P is given by P = \frac{\sqrt{S_1^2 + S_2^2 + S_3^2}}{S_0}, which ranges from 0 for completely (where S_1 = S_2 = S_3 = 0) to 1 for fully polarized light. This metric indicates the fraction of the total intensity that arises from the coherent, polarized portion of the beam, with $1 - P representing the unpolarized contribution. In the context of coherence theory, the Stokes parameters are directly related to the coherency matrix, a 2×2 Hermitian matrix \mathbf{J} formed from the correlation functions of the electric field components E_x and E_y: \mathbf{J} = \begin{pmatrix} \langle E_x E_x^* \rangle & \langle E_x E_y^* \rangle \\ \langle E_y E_x^* \rangle & \langle E_y E_y^* \rangle \end{pmatrix}, where the angle brackets denote time averages. The elements of \mathbf{J} yield the Stokes parameters via S_0 = J_{11} + J_{22}, S_1 = J_{11} - J_{22}, S_2 = 2\Re(J_{12}), and S_3 = 2\Im(J_{12}), providing a bridge between field correlations and observable intensities for partially coherent light. Stokes parameters are measured experimentally by passing the light through combinations of quarter-wave plates and linear polarizers, with intensities recorded at specific orientations to isolate each component; for instance, S_3 requires a quarter-wave plate to convert circular polarization differences into linear ones detectable by the polarizer. Optical transformations, such as passage through birefringent elements or scattering media, are described using Mueller matrices, which are 4×4 real matrices that map input Stokes vectors to output ones, enabling prediction of polarization changes in complex systems. A key application of Stokes parameters arises in analyzing natural light sources like , which becomes partially polarized due to single by atmospheric molecules, typically exhibiting a degree of polarization up to around 80% in the band 90° from , with the location of maximum polarization shifting from near the horizon (when is high) to near the (when is low on the horizon), as observed in clear skies. This partial facilitates of atmospheric properties and biological in animals, highlighting the parameters' utility in bridging classical with practical observations.

Polarization Entanglement in Quantum Optics

Polarization entanglement occurs when two or more share a such that their are correlated in a way that cannot be described by , leading to non-local correlations observable in joint measurements. This phenomenon is a cornerstone of , enabling the creation of maximally entangled states known as Bell states, such as the state |\Phi^+\rangle = \frac{1}{\sqrt{2}} (|HH\rangle + |VV\rangle), where |H\rangle and |V\rangle denote horizontal and vertical , respectively. These states were first realized experimentally using (SPDC) in the 1990s, marking a significant advancement in generating high-fidelity entangled photon pairs for tasks. SPDC involves the nonlinear optical process in a birefringent where a splits into a pair of lower- signal and idler photons, conserving and . In type-II phase-matching configurations, such as those using beta-barium borate (BBO) crystals, the signal and idler photons emerge with orthogonal polarizations—typically one horizontal and one vertical—naturally forming an entangled state upon proper alignment and compensation for walk-off effects. This method, pioneered in high-intensity sources, produces polarization-entangled pairs with fidelities exceeding 90% and has become the standard for laboratory-scale experiments since the mid-1990s. The nonlocality of polarization entanglement manifests in violations of Bell inequalities, which test the incompatibility of with local hidden variable theories. In landmark experiments from the early 1980s, and collaborators used entangled pairs from atomic cascades to measure polarization correlations, achieving a clear violation of the Clauser-Horne-Shimony-Holt ( with a value of S = 2.697 \pm 0.015, surpassing the classical bound of 2. Subsequent adaptations to SPDC sources in the 1990s confirmed these violations with polarization-entangled s, solidifying the empirical evidence for in optical systems. Polarization-entangled states are highly sensitive to decoherence, where interactions with the environment—such as in optical fibers or in free space—disrupt the phase coherence between the polarization components, reducing entanglement . For instance, propagation through dispersive media can introduce , leading to stochastic decoherence that scales with distance and wavelength mismatch, necessitating active compensation techniques like wave plates or feedback loops to preserve entanglement for practical quantum protocols. Recent studies show that atmospheric can degrade entanglement over propagation distances, often reducing visibility significantly without mitigation. As of 2025, polarization entanglement has found critical applications in satellite-based (QKD), where entangled photon pairs are transmitted over global distances to enable . The 2017 Micius satellite demonstrated the distribution of polarization-entangled photons over 1,200 km, achieving CHSH violations in space-to-ground links, and ongoing missions continue to refine this for practical networks, with recent entanglement-based QKD protocols viable up to 400 km in low-Earth orbit under daylight conditions.

Applications in Quantum Information Processing

Photon polarization plays a central role in (QKD), particularly through the protocol, where information is encoded in the polarization states of single photons. In this protocol, the sender () prepares photons in one of four polarization states: horizontal (H), vertical (V), or the diagonal states at 45° and 135°. randomly chooses between two bases—rectilinear (H/V) or diagonal—for each photon, while the receiver () measures in a randomly selected basis, discarding mismatched measurements post-sifting. The security arises from the and the disturbance caused by any eavesdropping attempt, which introduces detectable errors exceeding 25% in the quantum bit error rate (QBER), allowing to verify privacy amplification. In quantum computing, photon polarization serves as a qubit encoding in linear optical systems, enabling universal quantum gates through interferometric setups. The Knill-Laflamme-Milburn (KLM) scheme demonstrates how polarization-encoded qubits, combined with beam splitters and single-photon detectors, can implement deterministic two-qubit gates like the controlled-NOT (CNOT) with by probabilistically teleporting nonlinear interactions. This approach leverages post-selection on outcomes to overcome the limitations of linear optics, which cannot directly create entanglement, achieving fault-tolerant computation with success probabilities approaching 1 using ancillary photons. encoding reduces the required number of spatial modes compared to path-based schemes, facilitating scalable integration. Quantum teleportation utilizes polarization-entangled photon pairs to transfer an unknown polarization state from one photon to another without physical transmission of the carrier. In the protocol, the sender performs a Bell-state measurement on the input photon and one half of the entangled pair, classically communicating the result (two bits) to the receiver, who applies a corrective Pauli operator to reconstruct the state on the distant photon. Experimental demonstrations began in 1998, achieving fidelities above 0.7 for polarization states over laboratory distances using sources. These implementations have since scaled to network distances, enabling quantum repeaters and . In quantum metrology, enhances sensing precision beyond classical limits, as seen in ghost imaging protocols with entangled photons. Here, spatial and correlations between signal and idler photons from down-conversion allow reconstruction of an object's -dependent profile using only bucket detection on one arm and a reference scan on the other, achieving sub-shot-noise . This entanglement-based method surpasses classical ghost imaging by factors of up to √2 in , with applications in of birefringent materials. Despite these advances, challenges persist in deploying polarization-based systems, particularly high losses and state decoherence in optical fibers due to and environmental perturbations, limiting transmission to tens of kilometers without active stabilization. Free-space links mitigate these issues, offering polarization stability over atmospheric paths up to 100 km, as demonstrated in satellite-based QKD experiments. As of 2025, integrated photonic chips address scalability by enabling on-chip polarization rotation and manipulation with losses below 1 dB, using or platforms for compact quantum processors and networks.

References

  1. [1]
    Photon Polarization - Richard Fitzpatrick
    A polarization can be ascribed to each individual photon (ie, quantum of electromagnetic radiation) in a beam of light.
  2. [2]
    [PDF] 1 Photon polarization Masatsugu Sei Suzuki Department of Physics ...
    Sep 20, 2014 · Polarization. The electric and magnetic vectors associated with an electromagnetic wave are perpendicular to each other and to the direction ...
  3. [3]
    7.12: Polarized Light and Quantum Mechanics - Chemistry LibreTexts
    Jan 10, 2023 · The purpose of this tutorial is to use polarized light to illustrate one of quantum theory's deepest and most challenging concepts - the linear superposition.<|control11|><|separator|>
  4. [4]
    [PDF] Chapter 12 - Polarization - MIT OpenCourseWare
    We describe the varieties of possible polarization states of a plane wave: linear, circular and elliptical. iii. We describe “unpolarized light,” and ...
  5. [5]
    Classification of Polarization - HyperPhysics Concepts
    A plane electromagnetic wave is said to be linearly polarized. The transverse electric field wave is accompanied by a magnetic field wave as illustrated.Missing: definitions | Show results with:definitions
  6. [6]
    Science, Optics and You - Timeline - Augustin-Jean Fresnel
    Nov 13, 2015 · Using his inventions, Fresnel was the first to prove that the wave motion of light is transverse. He accomplished this task by polarizing light ...
  7. [7]
    (PDF) The Fresnel triprism and the circular polarization of light
    In 1822 Augustin Fresnel discovered the circular polarization of light with an experiment in which a plane polarized beam was resolved into its left- and right ...
  8. [8]
    [PDF] Linear Algebra for Describing Polarization and Polarizing Elements
    First formulated in 1941, Jones calculus utilizes a 2x1 vector to describe the polarization of light, and 2x2 matrices to describe the action of an element ...
  9. [9]
    The Poincaré Sphere - SPIE
    All linear polarization states lie on the equator and right and left circular polarization states are at the north and south poles, respectively. Elliptically ...<|control11|><|separator|>
  10. [10]
    Poincaré and his polarization sphere - Kahr - Wiley Online Library
    Sep 25, 2021 · In a text (1892) on light, Jules Henri Poincaré introduced a geometrical device for tracking the polarization of state of light interacting with matter.Missing: original | Show results with:original
  11. [11]
    The Jones vector as a spinor and its representation on the Poincaré ...
    Mar 19, 2013 · A SU(2) spinor that corresponds to a tangent vector to the Poincaré sphere representing the state of polarization and phase of the wave.
  12. [12]
    Revisiting Poincaré Sphere and Pauli Algebra in Polarization Optics
    Poincaré used this spherical representation for describing the transformation of optical polarization states (SOPs) under the action of various retarders and ...
  13. [13]
    The Poincaré-sphere approach to polarization: Formalism and new ...
    Nov 1, 2016 · The simplicity of the Poincaré sphere for visualizing polarization is that points of the same latitude have the same ellipticity, and points ...
  14. [14]
    Using the Poincaré Sphere to Represent the Polarization State
    However, a key benefit of the spherical representation is that it simplifies the math needed to calculate incremental changes in polarization state.
  15. [15]
    Understanding Polarized Light, Stokes Vectors, and the Poincaré ...
    The Poincaré sphere is a useful tool for visually describing polarization states. The Stokes vector is a four-element vector that fully defines the polarization ...Missing: stereographic | Show results with:stereographic
  16. [16]
  17. [17]
    [PDF] PLANE ELECTROMAGNETIC WAVES - UT Physics
    For r = 1 we get a linear polarization in the direction φ, for r = 0 we get a circular polarization, and for any other 0 <r< 1 we get an elliptic polarization.Missing: elliptical | Show results with:elliptical
  18. [18]
    XV. On the transfer of energy in the electromagnetic field - Journals
    Poynting J. H.. 1884XV. On the transfer of energy in the electromagnetic fieldPhil. Trans. R. Soc.175343–361http://doi.org/10.1098/rstl.1884.0016. Section.
  19. [19]
    27 Field Energy and Field Momentum - Feynman Lectures
    So the Poynting vector gives not only energy flow but, if you divide by c2, also the momentum density. The same result would come out of the other analysis we ...
  20. [20]
    [PDF] Unit 4-1: Electromagnetic Energy Density and the Poynting Vector
    Jan 4, 2022 · In this unit we extend the ideas of energy and momentum to electromagnetic fields, defining the energy density, energy current (Poynting vector) ...
  21. [21]
    16.3 Energy Carried by Electromagnetic Waves - UCF Pressbooks
    ... origin at t = 0 , the electric and magnetic fields obey the equations. E y ( x , t ) = E 0 cos ( k x − ω t ) B z ( x , t ) = B 0 cos ( k x − ω t ) . The energy ...16.3 Energy Carried By... · Learning Objectives · Problems
  22. [22]
    16.4 Momentum and Radiation Pressure - UCF Pressbooks
    The radiation pressure p rad applied by an electromagnetic wave on a perfectly absorbing surface turns out to be equal to the energy density of the wave:.
  23. [23]
    [PDF] Electromagnetic Momentum - Reed College
    “Electromagnetic momentum density and the Poynting vector in static fields,” F. S. Johnson, B. L. Cragin, and R. R. Hodges, Am. J. Phys. 62,. 33–41 (1994) ...
  24. [24]
    [PDF] Spin and Orbital Angular Momenta of Electromagnetic Waves ... - arXiv
    The intrinsic angular momentum of the pulse is the sum of its orbital (L) and spin (S) angular momenta, while the extrinsic part is given by rCM × p.
  25. [25]
    A decomposition of light's spin angular momentum density - Nature
    Jul 10, 2024 · The helicity density of a circularly polarised wave will be one (two) times the energy density and canonical spin will be one (two) times the ...
  26. [26]
    [PDF] Electromagnetic Angular Momentum - The University of Arizona
    Here. pp = SS(rr,tt) cc2. ⁄ is the EM linear momentum density,. SS = EE × HHis the Poynting vector, and cc is the speed of light in vacuum. Below we discuss two ...
  27. [27]
    Malus's Law - SPIE
    In 1808, using a calcite crystal, Malus discovered that natural incident light became polarized when it wasreflected by a glass surface.Missing: Etienne- Louis
  28. [28]
    A New Calculus for the Treatment of Optical Systems. IV.
    R. Clark Jones, "A New Calculus for the Treatment of Optical Systems. IV.," J. Opt. Soc. Am. 32, 486-493 (1942). Export Citation. BibTex; Endnote (RIS); HTML ...
  29. [29]
    Jones Calculus - SPIE
    An application of the Jones matrix calculus is to determine the intensity of an output beam when a rotating polarizer is placed between two crossed polarizers.
  30. [30]
    Polarizers - RP Photonics
    Absorptive filters can handle only quite limited optical powers (have a low optical damage threshold) because the absorbed power is converted to heat, and the ...
  31. [31]
    Some Aspects of the Development of Sheet Polarizers*
    ### Summary of Polaroid H-Sheet Polarizers
  32. [32]
    Principles of Birefringence | Nikon's MicroscopyU
    2. Birefringence (B) = |n e - no|. where n(e) and n(o) are the refractive indices experienced by the extraordinary and ordinary rays, respectively. This ...
  33. [33]
    Calcite - Birefringence - HyperPhysics
    This ray is called the "extraordinary ray". The indices of refraction for the o- and e-rays are 1.6584 and 1.4864 respectively.
  34. [34]
    Double Refraction | Harvard Natural Sciences Lecture Demonstrations
    A birefringent substance will split unpolarized light into two polarized rays with different refractive indices and different velocities.
  35. [35]
    [PDF] Birefringence
    Light traveling in any other direction through the crystal experiences two different refractive indices and is split into components that travel at different.
  36. [36]
    Phase retardation in birefringent material - vCalc
    Jul 24, 2020 · The following formula is used: Γ=2πΔnLλ0 Γ = 2 π Δ n L λ 0 , where: Δn Δ n = Change in the index of refraction; L L = thickness of the crystal ...Missing: delta = d / lambda n_o)
  37. [37]
    Recent Advances in Birefringence Studies at THz Frequencies
    May 9, 2013 · \varDelta \phi =\frac{{2\pi }}{\lambda }d\Delta n. This effect is commonly exploited for the fabrication of half or quarter wave plates ...Missing: formula | Show results with:formula
  38. [38]
    [PDF] Lecture 28 – Polarization of Light - Purdue Physics
    – Propagation through birefringent material. • Measuring polarization tells us ... – Jones calculus quantifies the phase evolution of the electric field ...
  39. [39]
  40. [40]
  41. [41]
    Wollaston Prisms - Evident Scientific
    A Wollaston prism is composed of two geometrically identical wedges of quartz or calcite (birefringent, or doubly-refracting materials), cut in a way that their ...Missing: history 1803
  42. [42]
    [PDF] arXiv:physics/0408069v2 [physics.optics] 7 Feb 2006
    expresses energy conservation of the light beam. This is just to be ... in an elegant way: given any linear lossless device with unitary Jones matrix ...
  43. [43]
    [PDF] Generalized polarization transformations with metasurfaces
    Nov 22, 2021 · Physically, a retarder or a waveplate can perform a unitary Jones matrix transformation. (b) A Hermitian transformation, H, satisfies H† = H ...<|separator|>
  44. [44]
    [PDF] Unitary and non-unitary operations on the Poincaré sphere ... - arXiv
    May 3, 2022 · In polarization optics unitary and non-unitary operations can be carried out by the Jones matrix. Z matrix is the 4 × 4 analogue of the Jones ...
  45. [45]
    Complete birefringence and Jones matrix characterization using ...
    We can then obtain the retarder matrix, as J R = J ⋅ J D − 1 . From this decomposition, the Jones matrix can be expressed as a product of homogeneous matrices ...
  46. [46]
    [PDF] Angle-to-retardance converter and universal polarization-state ...
    Jones matrices of retarders are unitary. ( †. 1. -. = T. T , the dagger indicating conjugate transpose) and depend on up to three free parameters, namely the ...
  47. [47]
    [PDF] Quantum Mechanics: Week 3 overview
    Physical realization: photon polarization. 2 dimensional Hilbert space H. Physical realization: electron spin. |ψi = cosθ|Hi + eiφ sinθ|V i. HV. ( 1. 0. ) ( 0.
  48. [48]
    Single Photon Interference | Harvard Natural Sciences Lecture ...
    The which-path marker consists of two, mutually perpendicular, polarizing filters. When either the vertical or the horizontal filter covers both slits, the ...
  49. [49]
  50. [50]
    [PDF] Lecture Notes for Ph219/CS219: Quantum Information Chapter 2
    1) to the photon polarization state. ... d) State and prove the result corresponding to (c) that applies to a n-part system with Hilbert space H1 ⊗ H2 ⊗···⊗Hn.
  51. [51]
    [PDF] Quantum Mechanics of Photon and Biphoton Experiments
    Jul 4, 2022 · We can also verify that the rotation operator is unitary. In the optics bag of tricks we have the half-wave plate. It is a birefringent device ...
  52. [52]
    [PDF] Theory for the beam splitter in quantum optics - arXiv
    Nov 7, 2022 · Abstract. The theory of the beam splitter (BS) in quantum optics is well developed and based on fairly simple mathematical and physical ...
  53. [53]
    [PDF] 1 The operator for photon polarization - bingweb
    Sep 8, 2025 · Here we discuss the Hermitian operator for photon polarization. These operators are derived from the projection operators.
  54. [54]
    [PDF] Lecture 9: Polarization in quantum optics.
    Now we will re-consider the description of polarization we used before. In quantum optics, every physical value corresponds to some operator.
  55. [55]
    Assessing the Polarization of a Quantum Field from Stokes ...
    We propose an operational degree of polarization in terms of the variance of the Stokes vector minimized over all the directions of the Poincaré sphere.Missing: paper | Show results with:paper
  56. [56]
    Quantum phase difference, phase measurements and Stokes ...
    Aug 7, 2025 · This admits very general scenarios, including especially the phase difference between two modes of the electromagnetic field.
  57. [57]
    Experimental Test of Error-Disturbance Uncertainty Relations by ...
    Jan 15, 2014 · In this Letter, we report the experimental test of the EDR for a single-photon polarization measurement using the weak-probe method. Our ...
  58. [58]
    On the Composition and Resolution of Streams of Polarized Light ...
    On the Composition and Resolution of Streams of Polarized Light from different Sources · George Gabriel Stokes; Book: Mathematical and Physical Papers; Online ...Missing: 1852 | Show results with:1852
  59. [59]
  60. [60]
    The Stokes Polarization Parameters - SPIE
    The first Stokes parameter S0 describes the total intensity of the optical beam; the second parameter S1 describes the preponderance of LHP light over LVP light ...
  61. [61]
    Stokes-Algebra Formalism - Optics Letters - Optica
    A unified treatment of the algebra of Stokes parameters and the coherency matrix is presented. Explicit formulas are given which relate the Jones and ...
  62. [62]
    [PDF] Measuring the Stokes polarization parameters
    (7). The Stokes parameters describe not only completely po- larized light but also unpolarized and partially polarized light as well. To describe these ...
  63. [63]
    Measurement of Stokes Parameters by Quarter-Wave Plate and ...
    The Stokes parameters (S0, S1, S2 and S3) need to be measured over a wide wavelength range. The Stokes parameters of monochromatic light can be measured by the ...
  64. [64]
    [PDF] Polarized light and the STOKES PARAMETERS - UVIC
    Its wavelength or frequency. Its wavelength depends upon the refractive index of the material in which it is travelling, whereas its frequency does not.
  65. [65]
    New High-Intensity Source of Polarization-Entangled Photon Pairs
    Dec 11, 1995 · We report on a high-intensity source of polarization-entangled photon pairs with high momentum definition.
  66. [66]
    Experimental Test of Bell's Inequalities Using Time-Varying Analyzers
    Each analyzer amounts to a polarizer which jumps between two orientations in a time short compared with the photon transit time. The results are in good ...
  67. [67]
    First Detailed Study of the Quantum Decoherence of Entangled ...
    The decoherence of propagating entangled optical photons, due to the continuous monitoring of the state by the environment, is well studied theoretically and ...
  68. [68]
    [2501.17130] Entanglement-based Quantum Key Distribution in the ...
    Jan 28, 2025 · We demonstrate that entanglement-based QKD is feasible under uplink daylight conditions in LEO satellites, but only up to a distance of 400 km.
  69. [69]
    Secure quantum key distribution with realistic devices
    May 26, 2020 · In the BB84 protocol, a sequence of single photons carrying qubit states is sent by Alice to Bob through a quantum channel. A schematic diagram ...
  70. [70]
    High-fidelity linear optical quantum computing with polarization ...
    We show that the KLM scheme [Knill, Laflamme, and Milburn, Nature 409, 46 (2001)] can be implemented using polarization encoding, thus reducing the number ...
  71. [71]
    Experimental quantum teleportation - Nature
    Dec 1, 1997 · During teleportation, an initial photon which carries the polarization that is to be transferred and one of a pair of entangled photons are ...
  72. [72]
    Fiber-coupled broadband quantum memory for polarization ... - Nature
    Oct 17, 2025 · Various near-term quantum networking applications will benefit from low-loss, fiber-coupled photonic quantum memory devices with high ...
  73. [73]
    FSO-QKD protocols under free-space losses and device imperfections
    May 13, 2024 · The main challenge of implementing QKD protocols over free-space link is due to atmospheric losses [31]. Other parameters such as timing, ...
  74. [74]
    Quantum photonics on a chip - AIP Publishing
    Jun 10, 2025 · Integrated quantum photonics uses photonic integrated circuits on a chip to control photonic quantum states for applications in quantum ...