Fact-checked by Grok 2 weeks ago

Optical medium

An optical medium is any material or substance through which electromagnetic waves, particularly , can propagate, characterized fundamentally by its , which quantifies the reduction in the compared to its velocity in a . These media interact with via , reemission, and processes at the level, leading to phenomena such as , , and that alter the wave's direction, speed, and . The behavior of light in an optical medium is governed by Maxwell's equations, with the medium's permittivity and permeability influencing the phase velocity (v_p = c/n, where c is the speed of light in vacuum and n is the refractive index) and group velocity of the propagating wave. In transparent media, light travels as plane waves with minimal absorption, while in absorbing or amplifying media, the complex refractive index (N = n + iκ) accounts for exponential decay or growth of the wave amplitude, where κ represents the extinction coefficient. Optical density, distinct from physical density, refers to the medium's tendency to impede light propagation due to repeated absorption and reemission by atoms, resulting in an effective speed less than c (approximately 3 × 10^8 m/s in vacuum). Optical media are classified into isotropic (uniform in all directions, e.g., or air, with n ≈ 1.5 for crown glass), anisotropic (direction-dependent index due to , exhibiting with ordinary index n_o and extraordinary index n_e, e.g., ), and inhomogeneous types (spatially varying index, e.g., graded-index fibers). Dispersion, where n varies with frequency (typically higher for shorter wavelengths), causes phenomena like in lenses and is described by models such as the . At interfaces between media, (n_1 sin θ_1 = n_2 sin θ_2) dictates , while determine and transmission coefficients, enabling applications from to .

Fundamentals

Definition and Scope

An optical medium is a through which electromagnetic in the optical range propagate, characterized by its interaction with via properties such as and permeability that influence wave speed and . These media encompass substances like dielectrics, crystals, and gases, where the response to incident arises from induced dipoles and collective motions, distinguishing them from non-optical that do not support such . The scope of optical media is confined to the optical portion of the , spanning wavelengths from approximately 100 nm in the to 1 mm in the , encompassing , visible, and radiation while excluding shorter wavelengths like X-rays and longer ones like microwaves or radio waves. Unlike a , where the is the constant c \approx 3 \times 10^8 m/s, optical media reduce this speed due to interactions with the material, with the serving as a key measure of this effect. The concept of the optical medium emerged in 19th-century , rooted in the wave theory of advanced by through his studies on and around 1816–1818, which demonstrated 's propagation through various substances as waves rather than particles. This framework built on earlier ideas but formalized the role of media in altering 's path and intensity, laying the groundwork for modern optical physics.

Wave Propagation Basics

In non-magnetic optical media, where the magnetic permeability \mu approximates the \mu_0, the propagation of electromagnetic waves is described by in the simplified form: \nabla \times \mathbf{E} = -\frac{\partial \mathbf{B}}{\partial t}, \quad \nabla \times \mathbf{H} = \mathbf{J} + \frac{\partial \mathbf{D}}{\partial t}, with constitutive relations \mathbf{D} = \epsilon \mathbf{E} and \mathbf{B} = \mu_0 \mathbf{H}, where \mathbf{J} denotes free currents (often negligible in dielectrics) and \epsilon is the of the medium. These equations yield the wave equation for the , \nabla^2 \mathbf{E} - \mu_0 \epsilon \frac{\partial^2 \mathbf{E}}{\partial t^2} = 0, enabling the propagation of transverse electromagnetic waves at speeds determined by the medium's properties. For monochromatic plane waves of the form \mathbf{E} = \mathbf{E_0} e^{i(\mathbf{k} \cdot \mathbf{r} - \omega t)}, the phase velocity v_p, which characterizes the speed of constant phase surfaces, is given by v_p = c / n, where c = 1/\sqrt{\mu_0 \epsilon_0} is the speed of light in vacuum and n = \sqrt{\epsilon / \epsilon_0} is the refractive index. This velocity is typically less than c in optical media, altering the wavelength while preserving the frequency \omega. In dispersive media, where n varies with frequency, the group velocity v_g = d\omega / dk becomes essential for signal propagation, representing the velocity of wave packets or pulses that carry information and energy. The distinction arises because v_g accounts for the envelope's motion in broadband signals, often differing from v_p and always subluminal in causal media. At interfaces between optical media, boundary conditions derived from Maxwell's equations ensure physical continuity: the tangential components of \mathbf{E} and the normal components of \mathbf{D} (absent free surface charges) remain continuous across the boundary. These conditions dictate how waves transmit and reflect, maintaining field integrity without singularities. Polarization plays a key role in propagation, with electromagnetic waves exhibiting linear polarization when the \mathbf{E}-field oscillates in a fixed direction, or circular polarization when it rotates uniformly, forming a helix. In birefringent media, which are anisotropic, the refractive index varies with polarization direction, causing orthogonally polarized components to propagate at different speeds and potentially altering the overall polarization state. Absorption within the medium can lead to gradual energy loss, exponentially decaying the wave amplitude over distance.

Physical Properties

Refractive Index

The refractive index n of an optical medium quantifies the reduction in the phase velocity of light propagating through it compared to vacuum, defined as n = \frac{c}{v}, where c is the speed of light in vacuum and v is the phase velocity in the medium. This property fundamentally governs light bending at boundaries between media, as described by Snell's law. For non-magnetic, non-absorbing media, n typically ranges from about 1.43 for fluorite to 2.42 for diamond at visible wavelengths, reflecting the medium's dielectric response to electromagnetic waves. The refractive index exhibits wavelength dependence, known as dispersion, expressed as n(\lambda), which causes different colors of to travel at slightly different speeds within the medium. This variation arises from the frequency-dependent of the medium's atoms or molecules. A widely used empirical model for this dispersion in transparent media is the , originally proposed by Wolfgang Sellmeier in 1871: n^2(\lambda) = 1 + \sum_i \frac{B_i \lambda^2}{\lambda^2 - C_i}, where \lambda is the in , and B_i and C_i are empirically fitted coefficients specific to the , often representing contributions from electronic resonances. This equation accurately predicts n(\lambda) over broad spectral ranges, such as from to near-infrared, for materials like fused silica where two or three terms suffice. In media that absorb light, the refractive index becomes complex, denoted \tilde{n} = n + i\kappa, where n is the real part influencing phase propagation and \kappa is the extinction coefficient that accounts for amplitude attenuation due to absorption. The imaginary part \kappa is related to the absorption coefficient \alpha by \alpha = \frac{4\pi \kappa}{\lambda}, linking it directly to energy loss mechanisms like electronic transitions. For weakly absorbing media, \kappa is small (e.g., $10^{-6} for high-purity in the visible), but it increases sharply near absorption bands. Refractive index is measured using techniques that exploit light deviation or path length changes. The minimum deviation method with a involves directing a monochromatic through the and measuring the angle of D, from which n = \frac{\sin((A + D)/2)}{\sin(A/2)}, where A is the apex angle; this yields accuracies of $10^{-4} or better for solids and liquids. Interferometric methods, such as the , determine n by observing fringe shifts when a sample is inserted into one arm, altering the by (n - 1)t, where t is the sample thickness; this is particularly precise for gases and thin films, achieving resolutions down to $10^{-6}. The depends on environmental factors like and , primarily through changes in and electronic structure. The thermo-optic coefficient \frac{dn}{dT} quantifies sensitivity, combining intrinsic polarizability shifts and effects; for example, in fused silica, \frac{dn}{dT} \approx 1.2 \times 10^{-5} \, \mathrm{K}^{-1} at 633 nm near . dependence, captured by \frac{dn}{dP}, arises similarly from variations, with negative values around −(1–2) × 10^{-6} ^{-1} for glasses and positive ~2 × 10^{-5} ^{-1} for , influencing applications in high-pressure optics.

Absorption, Dispersion, and Attenuation

In optical media, refers to the process by which is converted into or other forms of within the material, reducing the of the propagating . This is quantitatively described by Beer's law, which states that the transmitted I through a medium of thickness z is given by I = I_0 e^{-\alpha z}, where I_0 is the initial and \alpha is the coefficient. The coefficient \alpha is related to the imaginary part \kappa of the complex refractive index by \alpha = 4\pi \kappa / \lambda, where \lambda is the in ; this relation arises from the interaction of the electromagnetic with the material's response. Absorption in optical materials primarily stems from electronic transitions between energy bands, interactions with lattice vibrations (phonons), and the presence of impurities or defects that introduce additional energy levels. Electronic transitions dominate in the and visible regions, where photons excite electrons from to conduction bands, while phonon-assisted processes contribute to absorption through multiphonon interactions. Impurities, such as ions or hydroxyl groups in silica, create localized bands that can significantly increase losses even at trace concentrations. These mechanisms define transparency windows for materials; for example, fused silica exhibits low over the range of approximately 0.2 to 3.5 \mum, making it ideal for and UV applications, though losses rise sharply beyond these limits due to OH-related vibrations near 2.7 \mum and multiphonon above 3.5 \mum. Attenuation encompasses all mechanisms that diminish during , including and , and is expressed as the total coefficient \alpha_{\text{total}} = \alpha_{\text{abs}} + \alpha_{\text{scatter}}. In optical fibers, is commonly quantified in decibels per kilometer (dB/km), where a loss of 0.2 dB/km at 1550 nm in low-loss silica fibers allows over hundreds of kilometers with minimal . losses, primarily from density fluctuations, scale as $1/\lambda^4 and are more pronounced at shorter wavelengths, while contributes variably based on purity. Loss spectra for common media, such as silica (records approaching 0.14 dB/km near 1550 nm for silica-core and 0.09 dB/km for hollow-core as of 2025) and fluoride glasses (extending transparency to mid-infrared with ~0.1 dB/km at 2.5 \mum), serve as key figures of merit for selecting materials in and bulk applications. Dispersion in optical media describes the wavelength-dependent variation in the propagation speed of light, leading to temporal broadening of optical pulses. The (GVD) parameter \beta_2 = d^2\beta / d\omega^2, where \beta is the and \omega is the , quantifies second-order effects, but in fiber , the chromatic parameter D = d(1/v_g)/d\lambda (in ps/(nm·km)) is widely used, with positive values indicating normal (longer s travel faster) and negative values anomalous . Pulse broadening arises from this differential group delay, limiting high-bit-rate communications; for instance, in standard single-mode fibers, D \approx 17 ps/(nm·km) at 1550 nm causes significant distortion over long distances without compensation. comprises material , due to the wavelength dependence of the , and , arising from the modal structure in confined geometries like fibers, where the latter can be engineered to flatten the total D across the C-band (1530–1565 nm).

Types of Optical Media

Solid Optical Media

Solid optical media encompass a diverse class of rigid materials that facilitate the propagation, manipulation, and confinement of through their structured atomic arrangements. These materials are characterized by their high mechanical stability and ability to maintain precise , making them essential for various photonic applications. Unlike fluids or gases, solid optical media exhibit fixed geometries that enable engineered profiles and low scattering losses. Key categories include crystalline, amorphous, , and ceramic-based solids, each offering unique optical behaviors influenced by their molecular organization. Crystalline solids, such as quartz (SiO₂) and diamond, demonstrate well-ordered atomic lattices that result in predictable light interactions. Quartz, a common uniaxial crystal, has a refractive index of approximately 1.544 for the ordinary ray at 589 nm, enabling its use in birefringent devices due to the difference between ordinary and extraordinary indices. Diamond exhibits one of the highest refractive indices among natural solids at about 2.42 in the visible range, attributed to its dense carbon lattice, which minimizes absorption and maximizes transparency from ultraviolet to infrared wavelengths. In uniaxial crystals like calcite (CaCO₃), optical anisotropy arises from the directional dependence of the refractive index, with the ordinary index around 1.66 and the extraordinary index about 1.49 at 589 nm, leading to double refraction where light splits into two polarized beams along the optic axis. This birefringence stems from the crystal's trigonal symmetry, where carbonate groups align perpendicular to the optic axis, altering polarizability differently in orthogonal directions. Amorphous solids, lacking long-range order, provide isotropic with reduced compared to polycrystalline materials. Fused silica, an amorphous form of SiO₂, is renowned for its exceptionally low optical loss, typically below 0.2 dB/km in the near-infrared, due to its high purity and absence of grain boundaries. This material maintains transparency over a broad spectrum from to mid-infrared, with a of about 1.46, and exhibits minimal suitable for precision . Polymers like (PMMA) offer flexibility in fabrication and a of approximately 1.49 in the visible range, making them viable for lightweight optical components despite higher absorption than inorganic glasses. Semiconductor solids, such as (GaAs), combine electronic and optical functionalities through their bandgap structures. GaAs has a of around 3.4 in the and a direct bandgap of 1.43 eV at , which determines its transparency window starting above approximately 870 nm, where energies fall below the bandgap and drops sharply. This bandgap confines optical transmission to longer wavelengths, with the material's high index enabling efficient waveguiding in integrated photonic circuits. Doping in like GaAs adjusts carrier concentrations, thereby tuning the refractive index via the for dynamic optical control. Ceramics and composites extend solid optical media to multifunctional structures, often incorporating periodic arrangements for advanced light management. Yttrium aluminum (YAG, Y₃Al₅O₁₂) ceramics, fabricated as transparent polycrystalline materials, replicate single-crystal with inline transmission exceeding 80% from visible to near-infrared, owing to their cubic structure that minimizes losses. Photonic crystals, composite solids with periodic variations on the scale of light wavelengths, create bandgap effects analogous to electronic bandgaps in semiconductors, prohibiting propagation of specific frequencies and enabling phenomena like . These structures, typically formed from materials like or silica, rely on lattice periodicity to control dispersion. Fabrication techniques for solid optical media emphasize precise control over composition and structure to tailor refractive indices and minimize defects. Doping introduces impurities, such as rare-earth ions in YAG or aluminum in ZnO ceramics, to modify the and thus the by up to several percent, enhancing functionalities like . Thin films of optical solids are commonly deposited via (CVD), where precursor gases react on heated substrates to form uniform layers with thicknesses from nanometers to micrometers, ensuring low and controlled index gradients for antireflection coatings or waveguides. In solids, absorption generally arises from electronic transitions or phonons, but high-purity fabrication keeps it below 10⁻⁶ cm⁻¹ in the transmission bands for materials like fused silica.

Fluid and Gaseous Optical Media

Fluid and gaseous optical media encompass liquids and gases that transmit , characterized by their ability to flow and respond to environmental changes, unlike rigid solids. These media are crucial in applications ranging from to tunable photonic devices, where their refractive indices and properties can be modulated by external conditions such as , , or composition. Liquids serve as versatile optical media due to their relatively high refractive indices and tunable . , a ubiquitous liquid medium, exhibits a refractive index of approximately 1.33 in the , enabling phenomena like in aqueous environments. However, displays strong for wavelengths greater than 1 μm, primarily due to vibrational overtones of O-H bonds, which limits its use in optics. Oils, such as mineral or variants, offer lower in certain bands and refractive indices around 1.4–1.5, making them suitable for immersion objectives in . Electro-optic fluids, exemplified by Kerr liquids like , undergo refractive index changes under applied electric fields via the , allowing dynamic control in devices such as optical shutters. Gaseous optical media, with refractive indices close to unity, play a key role in atmospheric propagation and laser systems. Air at standard temperature and pressure (STP) has a refractive index of approximately 1.0003 for visible light, resulting from weak polarizability of its molecular components. Noble gases like helium and argon exhibit even lower indices (e.g., 1.000036 for helium), valued for their transparency and minimal dispersion in high-vacuum applications. In denser gases, pressure broadening occurs, where increased molecular collisions widen spectral lines and alter absorption profiles, impacting precision spectroscopy. Plasmas, as ionized gaseous media, display unique optical behavior distinct from neutral gases. For electromagnetic frequencies above the plasma frequency \omega_p = \sqrt{\frac{N e^2}{\epsilon_0 m}}, where N is the , e the , \epsilon_0 the , and m the , the n falls below 1, leading to anomalous dispersion and . This property underpins applications in plasma-based mirrors and fusion diagnostics. Mixtures and aerosols extend the functionality of fluid and gaseous media through composite effects. , an aerosol of droplets in air, causes significant via Mie , reducing visibility and altering in . Liquid crystals, as anisotropic fluids, enable tunable refractive indices by aligning molecules with , facilitating switchable displays and adaptive lenses. Environmental factors, such as , influence gaseous indices; increased content decreases air's refractive index by approximately 3.7 \times 10^{-8} per percent relative humidity at 20°C, affecting long-range communication. in atmospheric gases, while generally weak, contributes to mirages and formation over long paths.

Light-Matter Interactions

Reflection and Refraction

When light propagating in an optical medium encounters an interface with another medium of different refractive index, part of the wave is reflected while the remainder is refracted, assuming the incident angle is below the critical value for total reflection. The refracted ray bends toward the normal if entering a denser medium (higher refractive index) and away from the normal if entering a rarer medium, preserving the continuity of the wave's phase across the boundary. This linear behavior at planar interfaces underpins many optical phenomena and devices, distinct from bulk propagation effects. Snell's law quantifies the refraction, stating that the product of the refractive index and the sine of the angle of incidence (or refraction) is constant across the interface: n_1 \sin \theta_1 = n_2 \sin \theta_2, where n_1 and n_2 are the refractive indices of the incident and transmitting , respectively, and \theta_1 and \theta_2 are the corresponding angles measured from the normal. This relation, derived from the boundary conditions of electromagnetic waves or of least time, ensures the tangential component of the wave remains continuous. For typical optical like air (n \approx 1) and glass (n \approx 1.5), light bends noticeably at oblique incidence, enabling applications such as lenses and prisms. The fraction of light reflected at the interface is described by the Fresnel equations, which account for polarization. For s-polarized light (electric field perpendicular to the plane of incidence), the reflectance R_s is given by R_s = \left| \frac{n_1 \cos \theta_1 - n_2 \cos \theta_2}{n_1 \cos \theta_1 + n_2 \cos \theta_2} \right|^2, where the angles are related by Snell's law. This formula arises from matching the tangential electric and magnetic fields at the boundary for plane waves. A parallel form exists for p-polarized light (electric field in the plane of incidence), showing higher transmission for p-polarization at most angles. At normal incidence, the equations simplify to R = \left( \frac{n_1 - n_2}{n_1 + n_2} \right)^2, yielding about 4% reflection for air-glass interfaces. Total internal reflection occurs when light travels from a medium with higher refractive index (n_1 > n_2) and the incident angle exceeds the \theta_c = \sin^{-1}(n_2 / n_1), resulting in complete reflection with no transmitted propagating wave. At and beyond \theta_c, the refracted angle would exceed 90°, which is impossible for propagating , so the reflectance reaches 100% for both polarizations. However, a non-propagating evanescent wave penetrates into the second medium, decaying exponentially with distance z from the interface as e^{-\kappa z}, where \kappa = \frac{2\pi}{\lambda} \sqrt{n_1^2 \sin^2 \theta_1 - n_2^2} and \lambda is the in ; this field carries no net across the boundary but enables near-field effects. For an air-glass interface at 45° incidence, the evanescent field drops to $1/e of its value within roughly $0.45 \lambda. Brewster's angle provides a special case where reflection vanishes for p-polarized light, occurring at \theta_B = \tan^{-1}(n_2 / n_1), such that the incident and refracted rays are perpendicular (\theta_1 + \theta_2 = 90^\circ). At this angle, derived from the p-polarization Fresnel equation setting the reflection coefficient to zero, the reflected ray would align with the dipole oscillation direction in the second medium, preventing radiation back into the first medium. For glass in air, \theta_B \approx 56^\circ, making it useful for polarizing beams by selecting p-polarization. To minimize unwanted reflection at interfaces, anti-reflection coatings exploit , with a simple single-layer design using a quarter-wave thickness (d = \lambda / (4 n_c), where n_c is the coating's and \lambda is the design ) and n_c = \sqrt{n_1 n_2} for normal incidence. This creates two reflections of equal but opposite at the air-coating and coating- interfaces, leading to destructive and near-zero at the design . For a (n_2 \approx 1.5) in air (n_1 = 1), a with n_c \approx 1.22 (e.g., ) reduces from 4% to under 1%. Such coatings are limited to narrow bandwidths and angles but are essential in lenses and solar cells.

Nonlinear Effects

In optical media, nonlinear effects arise when the response of the material to becomes dependent on the of the optical , deviating from the valid at low intensities. These effects stem from higher-order terms in the expansion, characterized by nonlinear susceptibilities such as \chi^{(2)} and \chi^{(3)}, enabling phenomena like frequency conversion and self-modulation that are crucial for advanced photonic devices. The , a third-order nonlinear process, induces an intensity-dependent given by n = n_0 + n_2 I, where n_0 is the linear refractive index, n_2 is the nonlinear refractive index coefficient, and I is the optical intensity. This leads to , where the phase of a propagating pulse shifts proportionally to its instantaneous intensity, broadening the pulse spectrum in media like optical fibers. Second-harmonic generation (SHG) is a second-order nonlinear process in non-centrosymmetric media, described by the polarization P = \epsilon_0 \chi^{(1)} E + \epsilon_0 \chi^{(2)} E^2 + \cdots, where the quadratic term generates a polarization oscillating at twice the frequency of the input field E. Efficient SHG requires matching to prevent destructive , often achieved through in crystals, where the and refractive indices align the velocities of the fundamental and harmonic waves. Two-photon absorption represents a nonlinear loss mechanism arising from the imaginary part of the third-order susceptibility \chi^{(3)}, enabling simultaneous absorption of two photons whose combined energy matches an electronic transition. This process, prominent in materials with suitable bandgaps, results in intensity-dependent attenuation and is quantified by the two-photon absorption coefficient, scaling with the square of the intensity. Optical solitons emerge in fibers when the Kerr-induced balances , maintaining pulse shape over long distances via solutions to the . These fundamental solitons, first theoretically predicted in the context of fiber propagation, enable distortion-free transmission at high powers. Key materials for these effects include (LiNbO₃), prized for its large \chi^{(2)} enabling efficient SHG and electro-optic modulation, and chalcogenide glasses, which exhibit high n_2 values due to their soft lattices, ideal for Kerr-based applications in the .

Applications and Uses

In Imaging and Displays

Optical media play a crucial role in systems through lenses designed to minimize aberrations, particularly , which causes different wavelengths of to at varying points. Traditional camera often employ achromatic doublets, consisting of a converging lens made from low-dispersion glass paired with a diverging lens of high-dispersion , to correct this issue by aligning focal points for and wavelengths. This combination leverages the differing and dispersions of the glasses— glass typically has a refractive index around 1.52 with a high (indicating low dispersion), while exceeds 1.62 with a low (higher dispersion)—to achieve sharper images across the without significant color fringing. In display technologies, liquid crystal displays (LCDs) utilize s as an optical medium to control light , enabling pixel-level modulation of transmitted light. Liquid crystals, nematic in structure, rotate the polarization plane of incident light when aligned at 45 degrees to crossed polarizers, producing bright states; applying an realigns the molecules parallel to the polarizers, blocking light for dark states. This manipulation, combined with ing, allows for high-contrast images with energy-efficient operation. Organic light-emitting diode (OLED) displays, in contrast, rely on thin organic layers as the emissive optical medium, where occurs as electrons and holes recombine in materials like small-molecule emitters or polymers, emitting light directly without a backlight. The of these layers, including refractive indices around 1.7–1.8, influence cavity effects that enhance color purity and efficiency through constructive . Microscopy benefits from immersion oils as optical media to bridge refractive index mismatches between the objective lens and specimen, boosting resolution beyond the diffraction limit of air (NA ≈ 1.0). Oils with refractive indices near 1.5, matching common glass coverslips, enable numerical apertures up to 1.4–1.5, improving lateral resolution by a factor of approximately 1.5 compared to dry objectives by reducing spherical aberration and increasing light collection. This index matching minimizes light scattering at interfaces, allowing finer details in biological samples to be resolved. Holography employs photopolymers as versatile recording media to capture patterns for three-dimensional image reconstruction. These self-developing materials, often acrylamide-based with photoinitiators, undergo upon exposure to laser light, creating refractive index modulations (Δn up to 0.01) that diffract readout light to reconstruct the hologram without wet processing. Photopolymers offer high (over 5000 lines/mm) and sensitivity, making them suitable for volume holograms in security features and . Metamaterials with negative refractive indices enable superlenses for subwavelength imaging, overcoming the diffraction limit of conventional . These engineered structures, composed of subwavelength resonators like split-ring arrays, exhibit effective negative permittivity and permeability, allowing evanescent waves to amplify rather than decay, thus resolving features smaller than λ/2—demonstrated experimentally at frequencies with resolutions down to λ/6. Such media facilitate applications in and biomedical imaging by focusing light to spots below the standard limit.

In Fiber Optics and Communications

Optical fibers serve as the primary optical medium for high-speed data in , enabling the of signals over vast distances with minimal loss. These fibers typically consist of a silica surrounded by a cladding layer, where the refractive index difference (Δn) between the core and cladding is approximately 0.01, facilitating and guiding the along the fiber. Single-mode fibers, with a core of about 8-10 micrometers, support only one spatial of , minimizing intermodal and allowing over hundreds of kilometers at high data rates, making them ideal for long-haul networks. In contrast, multimode fibers feature larger cores (50-62.5 micrometers) that accommodate multiple modes, suitable for shorter distances in local area networks where easier coupling is prioritized over . A key advantage of silica-based optical fibers in communications is their attenuation minimum around 1550 nm, where losses can be as low as 0.17 /km, enabling signals to travel over 1000 km with periodic amplification. This wavelength aligns with the C-band (1530-1565 nm), the standard for telecom systems, and is amplified efficiently by erbium-doped fiber amplifiers (EDFAs), which provide up to 30 gain with a noise figure of about 5 through stimulated emission in erbium ions. EDFAs operate seamlessly in this low-loss window, supporting dense wavelength division multiplexing (DWDM) by amplifying multiple channels simultaneously without converting signals to electrical form. Wavelength division multiplexing (WDM) further enhances capacity by transmitting multiple optical signals at distinct wavelengths—up to 160 channels spaced 12.5-100 GHz apart—over a single , achieving aggregate exceeding 100 Gbit/s per channel. In dispersion-managed links, alternating segments of positive and negative dispersion compensate for chromatic dispersion effects, which can broaden pulses in high-bit-rate WDM systems around 1550 nm. For applications requiring reduced nonlinear distortions, photonic bandgap fibers with hollow cores guide light primarily through air via periodic microstructures, exhibiting significantly lower nonlinearity than conventional silica-core fibers due to minimized light-matter interaction. Looking ahead, plastic optical fibers (POFs) are gaining traction for short-range communications, such as in-vehicle networks and home LANs, offering bandwidths up to 1 Gbit/s over distances under 100 meters with low-cost installation. Additionally, specialized optical fibers are emerging as media for quantum communication, supporting entanglement distribution over 250 km networks by preserving coherence in low-loss silica structures.

References

  1. [1]
    [PDF] Physics of Light and Optics
    Topics are addressed from a physics perspective and include the propagation of light in matter, reflection and transmission at boundaries, polarization effects,.
  2. [2]
    Optical Density and Light Speed - The Physics Classroom
    The mechanism by which a light wave is transported through a medium occurs in a manner that is similar to the way that any other wave is transported - by ...Missing: definition | Show results with:definition
  3. [3]
    Optical radiation - Bundesamt für Strahlenschutz
    Optical radiation is part of the electromagnetic spectrum with a wavelength range from 100 nm to 1 mm. The sun is the natural source of optical radiation.Missing: physics | Show results with:physics
  4. [4]
    July 1816: Fresnel's Evidence for the Wave Theory of Light
    In July 1816, Fresnel published preliminary results on diffraction, aiming to develop a full theory, building on Young's work and embracing the wave theory.
  5. [5]
    [PDF] Chapter 2 Classical Electromagnetism and Optics
    For dielectric non magnetic media, which we often encounter in optics, with ... MAXWELL'S EQUATIONS OF ISOTROPIC MEDIA. The polarization in media with a ...
  6. [6]
    [PDF] Chapter 6 Maxwell's Equations for Electromagnetic Waves
    These are measures of the ability of the electric and magnetic fields to. “permeate” the medium; if is increased, then a larger electric field exists within the.Missing: non- | Show results with:non-
  7. [7]
    Phase and group velocities
    In a vacuum, the wave clearly moves with velocity c = ω/k. Now, if that wave passes into a medium with refractive index n, then some of its properties may ...
  8. [8]
    [PDF] Group Velocity | Kirk T. McDonald
    Dec 4, 1996 · = ω/k but velocity of propagation of vibrating motion = dω/dk. ... example, B = 0 for an electromagnetic wave while A = 0 for the granular systems ...
  9. [9]
    [PDF] Lecture 11: Wavepackets and dispersion
    vp = vp(kc) is the phase velocity at kc and vg = vg(kc) is called the group velocity group velocity vg(k)= dω(k) dk. (25). In general, both the phase and group ...
  10. [10]
    [PDF] Electromagnetic Boundary Conditions - Rutgers Physics
    a, b ) The normal components of D and B, and c,d) The tangential components of E and H are continuous across an interface between two dissimilar materials.
  11. [11]
    [PDF] Polarization
    Polarization includes linear, where the direction is fixed, circular, where the direction rotates, and elliptical, a general case where both magnitude and ...
  12. [12]
    Polarization of Light - StatPearls - NCBI Bookshelf
    May 29, 2023 · Birefringence refers to the phenomenon where a material causes the polarization state of light to change as it passes through. This occurs ...
  13. [13]
    [PDF] Light Absorption (and Optical Losses) - DSpace@MIT
    Calculate reflectance and non-absorption optical losses of a solar cell. • Calculate reflection of an intereface (semi-infinite).
  14. [14]
    31 The Origin of the Refractive Index - Feynman Lectures
    It is this advance in phase which is meant when we say that the “phase velocity” or velocity of the nodes is greater than c. In Fig. 31–4 we give a ...Missing: optical | Show results with:optical
  15. [15]
    [PDF] TIE-29: Refractive Index and Dispersion
    Refractive index (n) is the ratio of light speed in vacuum to a medium, measured relative to air. Dispersion is the change of refractive index with wavelength.
  16. [16]
    Science, Optics, and You: Light and Color - Refraction of Light
    Nov 13, 2015 · Refractive index is defined as the relative speed at which light moves through a material with respect to its speed in a vacuum. By convention, ...
  17. [17]
    Refraction, Polarisation and Dispersion of Light Waves
    The refractive index of a material is defined as the ratio of the speed of light in a vacuum to the speed of light in the material. It tells us how much slower ...
  18. [18]
    Fitting refractive-index data with the Sellmeier dispersion formula
    This paper describes a method for the fitting of optical index data to the Sellmeier dispersion formula. The work described began in 1964. It was shown in ...
  19. [19]
    Sellmeier circuits: A unifying view on optical and plasma dispersion ...
    Jul 15, 2020 · The Sellmeier equation, ubiquitous in optics for characterizing an optical material's refractive index (n) as a function of wavelength (λ) ...
  20. [20]
  21. [21]
    [PDF] Refractive Index. - Manx Precision Optics
    Together with the index of refraction it constitutes the imaginary part of the “complex refractive index” N=n+ik where i is the imaginary unit. In optics ...<|separator|>
  22. [22]
    [PDF] Section 19 Dispersing Prisms
    At minimum deviation, the ray path through a dispersing prism is symmetric . The ray is bent an equal amount at each surface.
  23. [23]
    [PDF] 1 Topics Minimum deviation of angle in prism Masatsugu Sei Suzuki ...
    To measure the index of refraction n(λ) for a specified wavelength λ, one use a goniometer. By determining the magnitude of the angle between the two sides of ...
  24. [24]
    [PDF] Measuring the Refractive Index of Infrared Materials by Dual ...
    Each method uses interferometry as the means for their measurement of the refractive index. The first of these nondestructive methods uses a Michelson ...
  25. [25]
    [PDF] Determining the refractive index of liquids using a modified ... - LOUIS
    A Michelson interferometer was used in this experiment to measure the refractive index values. This chapter describes the experimental setup and the ...
  26. [26]
    [PDF] Infrared Refractive Index and Thermo-optic Coefficient Measurement ...
    The simplest expression of temperature dependence is the derivative of refractive index with respect to temperature, i.e., the thermo-optic coefficient.
  27. [27]
    [PDF] Temperature-dependent refractive index of CaF2 and Infralsil301
    For CaF2, we report absolute refractive index and thermo-optic coefficient. (dn/dT) at temperatures ranging fiom 25 to 300 K at wavelengths from 0.4 to 5.6 pm, ...
  28. [28]
    [PDF] The effect of temperature and pressure on the refractive index of ...
    The dependence of refractive index, n, upon tem- perature, T, is important in optical design where systems must operate at inconstant temperature or at.
  29. [29]
    [PDF] Measured relationship between thermodynamic pressure and ...
    cell length, and the change in refractive index as a function of pressure dn dp. Lastly, we note that an alternate cell design34 can poten- tially reduce ...
  30. [30]
    Correction to the Beer-Lambert-Bouguer law for optical absorption
    The Beer–Lambert–Bouguer absorption law, known as Beer's law for absorption in an optical medium, is precise only at power densities lower than a few kW.
  31. [31]
    Conditions for admittance-matched tunneling through symmetric ...
    Jul 5, 2012 · ... optical absorption coefficient (αm = 4πκm/λ) for Ag. Optical constants of Ag were modeled using the Lorentz-Drude expressions provided by ...
  32. [32]
    Absorption – light, processes, linear and nonlinear ... - RP Photonics
    There can be ionic impurities which introduce absorption based on electronic transitions. · Similarly, semiconductor nanocrystals can lead to absorption, in that ...
  33. [33]
    Optical constants of silica glass from extreme ultraviolet to far ...
    Practically, silica glass is transparent from 200 nm up to m 3.5 – 4.0 μ m . In the extreme UV (for wavelengths below ...Fig. 1 · Fig. 2 · Fig. 3 · Fig. 4<|control11|><|separator|>
  34. [34]
    Optical Fiber Loss and Attenuation - Fosco Connect
    The attenuation of an optical fiber measures the amount of light lost between input and output. Total attenuation is the sum of all losses.
  35. [35]
    Broadband optical fibre with an attenuation lower than 0.1 decibel ...
    Sep 1, 2025 · Simultaneously, optical fibres also presented an ultralow level of attenuation of around 0.15 dB km−1, which remained approximately constant ...
  36. [36]
    Group Velocity Dispersion - RP Photonics
    In optical fiber communications, the dispersion parameter D λ , with units of ps/(nm·km), is more common; it is related to β 2 but has the opposite sign.
  37. [37]
    Fresnel Equations - The University of Arizona
    Developed in the years 1821-1823, the Fresnel equations[1] describe the amplitude of transmitted and reflected light at the boundary between two materials.
  38. [38]
    1.4 Total Internal Reflection - University Physics Volume 3 | OpenStax
    Sep 29, 2016 · Total internal reflection occurs for any incident angle greater than the critical angle θ c θ c , and it can only occur when the second medium ...Missing: source | Show results with:source
  39. [39]
    None
    ### Summary of Brewster's Angle from Lecture 15
  40. [40]
    Anti-reflection Coatings - RP Photonics
    In the simplest case, an anti-reflection thin-film coating designed for normal incidence consists of a single quarter-wave layer of a material the refractive ...What are Anti-reflection... · Single-layer Anti-reflection... · Coatings with Strongly...
  41. [41]
    Kerr effect - RP Photonics
    The Kerr effect is a nonlinear optical effect which can occur when light propagates in crystals and glasses, but also in other media such as gases.
  42. [42]
    A tutorial on fiber Kerr nonlinearity effect and its compensation in ...
    Nov 22, 2021 · The Kerr effect induces a power-dependent nonlinear distortion for the signal propagating in the optical fiber. The detrimental effects of the ...
  43. [43]
    Birefringent Phase Matching – nonlinear frequency conversion
    Birefringent phase matching is a technique for achieving phase matching of a nonlinear process by exploiting the birefringence of a nonlinear crystal.
  44. [44]
    Two-photon Absorption – TPA
    Two-photon absorption is a nonlinear absorption process where two photons are simultaneously absorbed. This can occur at high optical intensities.
  45. [45]
    Two-photon absorption: an overview of measurements and principles
    Two-photon absorption (2PA) is the process by which a molecule or material absorbs a pair of photons, the sum of whose energy equals the transition energy. The ...
  46. [46]
    [PDF] Solitons in optical fibers
    The discovery of Optical Solitons dates back to 1971 when Zhakarov and. Sabat solved in 1971 the nonlinear Schrodinger (NLS) equation with the inverse ...
  47. [47]
    Recent development in integrated Lithium niobate photonics
    In conclusion, LNOI has both second-order nonlinear properties of monocrystalline materials and higher optical integration, providing fascinating platform for ...
  48. [48]
    9.6 Aberrations | TEKS Guide
    For example an achromatic doublet consisting of a converging lens made of crown glass and a diverging lens made of flint glass in contact can dramatically ...
  49. [49]
    [PDF] EVALUATION OF CORRECTION METHODS OF CHROMATIC ...
    Aug 25, 2012 · The most common type is an achromatic doublet composed of two elements made of crown glass and flint glass.
  50. [50]
    Common Optical Defects in Lens Systems (Aberrations)
    Sep 10, 2018 · Additional refinement of an optical system with even more sophisticated glass formulas and shapes can reduce chromatic aberration even further.
  51. [51]
    Liquid Crystal Displays - MIT Media Lab
    When linearly polarized light passes through, the optical activity of the material causes the polarization of the light to rotate by a certain angle. The ...
  52. [52]
    Top-emitting organic light-emitting diodes - Optica Publishing Group
    The properties of OLEDs, such as area-emission, high electron to photon conversion efficiency, wide viewing angle, low operating voltage, fast switching, vivid ...
  53. [53]
    Optical design for improving optical properties of top-emitting ...
    Mar 18, 2013 · The most commonly used structure for OLED is anode/organic layers/cathode. The organic layers are generally composed of the hole transport ...
  54. [54]
    Microscopy Basics | Objectives - Zeiss Campus
    A gain in resolution by a factor of about 1.5 is attained when immersion oil is substituted for air as the imaging medium. Finally, the last but perhaps ...
  55. [55]
    Impact of immersion oils and mounting media on the confocal ... - NIH
    Our data emphasize that to obtain the best optical resolution, the refractive index of immersion oil should match that of the glass through which the sample ...
  56. [56]
    Holographic Recording with Photopolymers*
    Phase holograms have been recorded using several photopolymers. These photopolymer recording systems are self-developing and require only uv-light fixing.
  57. [57]
    Holographic Photopolymer Material with High Dynamic Range (Δn ...
    Aug 26, 2020 · A high-performance holographic recording medium was developed based on a unique combination of photoinitiated thiol–ene click chemistry and functional, linear ...
  58. [58]
    Subwavelength resolution with a negative-index metamaterial ...
    Jun 19, 2007 · Negative-index metamaterials are candidates for imaging objects with sizes smaller than a half-wavelength.
  59. [59]
    Superlens imaging theory for anisotropic nanostructured ...
    For wavelength larger than that of the longitudinal SPR, these media are negative index metamaterials and can be used for superlens imaging 4, 13 in the ...
  60. [60]
    Refractive-index profiling of single-mode optical fibers and performs
    The maximum index difference between core and cladding is about 0.01. The index profile of the fiber pulled from this preform is shown in Fig. 4(b) .
  61. [61]
    applications, fiber optics, single-mode and multimode, polarization ...
    Single-mode fibers usually have a relatively small core (with a diameter of only a few micrometers) and can guide only a single spatial mode (disregarding the ...
  62. [62]
    100 Million Zoom Sessions Over a Single Optical Fiber
    With a pure silica core, its attenuation is specified at no more than 0.17 dB/km at the 1550-nanometer minimum-loss wavelength, close to the theoretical limit.
  63. [63]
    Fiber-based phase-sensitive optical amplifiers and their applications
    Erbium-doped fiber amplifier. FOPA. Fiber-optic parametric ... optical communication systems operating in the low-loss attenuation window around 1550 nm.
  64. [64]
    Wavelength Division Multiplexing - RP Photonics
    Wavelength division multiplexing is a kind of frequency division multiplexing – a technique where optical signals with different wavelengths are combined.
  65. [65]
    Comparison of Nonlinear Properties in Silica-Core and Hollow-Core ...
    May 9, 2024 · We experimentally evaluate whether the hollow-core photonic bandgap fiber has extremely low nonlinearity compared with silica-core optical fibers.
  66. [66]
    (PDF) Plastic optical fibers branch out - ResearchGate
    Aug 9, 2025 · POFs are in high demand in various high-volume applications for short-distance data communication, especially in home networks and in-vehicle ...
  67. [67]
    Quantum communication across a 250-kilometre optical-fibre network
    Apr 23, 2025 · Quantum communication over long distances can be achieved by exploiting a property of light called coherence.