Fact-checked by Grok 2 weeks ago

Optical properties

Optical properties of materials encompass the ways in which substances interact with , particularly in the , visible, and regions of the , through processes including , , , , and . These properties arise from the material's and structure, influencing how propagates, scatters, or is absorbed, and are fundamentally described by the \tilde{n} = n + i k, where n is the real part governing and k the imaginary part related to . Key measurements of optical properties include , which quantifies the fraction of incident reflected; , the fraction transmitted; , the fraction absorbed; emittance, related to ; and the index of refraction, which determines bending of at interfaces. These quantities often depend on factors such as , , angle of incidence, and temperature, and are linked to the material's dielectric function \epsilon(\omega), which connects optical behavior to and conductivity. In solids, interactions occur via mechanisms like electronic (dominant in dielectrics for visible ), vibrational modes (in ), and free electron responses (in metals, leading to high reflectivity below the plasma frequency). Optical properties are crucial for applications in , , and , enabling technologies such as lenses, solar cells, displays, and lasers, where precise control over light-matter interactions is essential. Techniques like , , and are used to characterize these properties, providing insights into material and without . Ongoing research explores nanostructured materials to engineer novel optical behaviors, such as enhanced absorption or .

Fundamental Concepts

Definition and Scope

Optical properties describe the interaction of materials with , particularly in the visible, , and spectra, manifesting through , , , and of . These properties arise from the structure of the material and dictate how is altered upon encountering , forming the basis for phenomena observable in everyday and scientific contexts. The scope of optical properties includes distinctions between linear and nonlinear responses, as well as isotropic and anisotropic behaviors. Linear optical properties characterize responses proportional to the incident light intensity, governing conventional light-matter interactions at low intensities. Nonlinear optical properties, in contrast, occur at high intensities where the response scales with higher powers of the field, enabling advanced effects like harmonic generation. Materials are further categorized as isotropic, with uniform properties independent of light propagation direction, or anisotropic, where directional variations stem from structural asymmetry, such as in crystals. A key aspect of optical properties is their dependence on , which influences material characteristics like , color, and luster. results from low across the visible range, permitting unimpeded passage; color emerges from wavelength-selective or tied to transitions; and luster reflects high reflectivity that imparts a glossy sheen. The stands as a fundamental measure of 's speed reduction in the material. These attributes are pivotal in and , driving applications in lenses and displays.

Historical Overview

The study of optical properties began with early observations of light's behavior in the . In 1666, conducted pivotal experiments using prisms to demonstrate and the of white light into a of colors, challenging prevailing views and laying the groundwork for understanding light as composed of distinct rays with varying refractive behaviors. Shortly thereafter, in 1678, proposed his wave theory of light, which explained as the of wavefronts through an , providing a mechanistic basis for light's interaction with matter that contrasted with Newton's corpuscular model. The 19th century saw significant advances in modeling light-matter interactions at interfaces. developed his equations between 1821 and 1823, deriving the amplitude reflection and transmission coefficients for light at dielectric boundaries, which quantitatively described and based on and . In 1845, discovered magneto-optical effects, observing that a could rotate the plane of polarization in transparent media like , revealing the influence of external fields on optical properties and foreshadowing electrodynamics' unification. In the early 20th century, informed classical models of optical responses in materials. Paul Drude's 1900 model treated metals as a gas of free electrons, predicting their high reflectivity and absorption in the visible range due to plasma-like oscillations, which became foundational for understanding metallic optical properties. The invention of the by in 1960 revolutionized the field, providing intense coherent light sources that enabled the exploration of nonlinear optical properties, where light intensity alters material responses beyond linear approximations. More recently, theoretical predictions have driven innovations in engineered materials. In 1968, Victor Veselago theorized substances with simultaneously negative permittivity and permeability, leading to where bends oppositely to conventional media, a concept initially unrealized until experimental demonstrations with metamaterials in 2001 using split-ring resonators and wire arrays. These milestones trace the evolution from empirical observations to quantum and engineered paradigms in optical properties.

Linear Optical Properties

Refractive Index and Dispersion

The refractive index n of a medium is defined as the ratio of the speed of light in vacuum c to the phase velocity v of light in that medium, quantifying how much slower light propagates through the material compared to free space. In general, n is a complex quantity expressed as n = n_r + i \kappa, where the real part n_r governs refraction and phase velocity, while the imaginary part \kappa relates to absorption, with details on absorption covered in subsequent sections on transmission properties. Dispersion refers to the wavelength dependence of the , arising primarily from electronic resonances in the material that cause n_r to vary with the light's . In dispersion, typical for transparent materials far from bands, n_r decreases with increasing (or equivalently, increases with ), leading to shorter s (e.g., ) refracting more than longer ones (e.g., red light), as observed in prisms. Conversely, anomalous occurs near resonances, where dn_r / d\lambda > 0, reversing the usual ordering and causing longer s to refract more strongly, though this region is narrow and often overlaps with high . Empirical formulas model this dispersion effectively for many dielectrics. The Cauchy dispersion formula, proposed in 1836, approximates n(\lambda) for visible wavelengths in low-dispersion media as n(\lambda) = A + \frac{B}{\lambda^2} + \frac{C}{\lambda^4}, where A, B, and C are fitted coefficients, with higher-order terms capturing finer variations but often negligible. For broader wavelength ranges, including near ultraviolet and infrared, the provides a more accurate representation, derived in 1871 and based on Lorentz oscillator contributions: n^2(\lambda) = 1 + \sum_i \frac{B_i \lambda^2}{\lambda^2 - C_i}, where B_i and C_i are material-specific parameters related to resonance strengths and squared resonance wavelengths, respectively, enabling precise from sparse measurements. In anisotropic materials, birefringence manifests as direction-dependent refractive indices, a linear optical effect stemming from the material's crystal symmetry. Uniaxial materials, such as or , possess one optic axis along which light experiences a single index, with perpendicular polarizations seeing (n_o) and (n_e) indices, yielding birefringence \Delta n = |n_e - n_o|. Biaxial materials, like or , have two optic axes and three distinct principal indices, complicating polarization but still characterized linearly by pairwise differences in indices without intensity dependence.

Absorption and Transmission

Absorption in optical refers to the process by which incident is dissipated as or other forms of , reducing the of the transmitted beam without altering its phase significantly. The primary quantitative measure of this phenomenon is the , denoted as α, which describes the fractional decrease in per unit distance traveled through the . According to the Beer-Lambert law, the transmitted I through a of thickness d is given by I = I_0 e^{-\alpha d}, where I_0 is the incident ; this law assumes monochromatic and a homogeneous, non-scattering medium. The absorption coefficient α is intrinsically linked to the material's , \tilde{n} = n + i\kappa, where n is the real part (related to ) and κ is the imaginary part (). Specifically, \alpha = \frac{4\pi \kappa}{\lambda}, with λ denoting the in ; this relation arises from the of the amplitude within the material, E(z) = E_0 e^{i(\omega t - \tilde{n} k_0 z)}, where the imaginary component leads to . Light occurs through various microscopic mechanisms depending on the material and . In semiconductors, transitions dominate, particularly near the bandgap E_g: when a photon's h\nu equals or exceeds E_g, it excites an from the to the conduction , generating an electron-hole pair and dissipating the excess as via interactions. In molecular materials, vibrational prevails in the (IR) range, where photon matches the quantized vibrational modes of chemical bonds, such as or bending; for instance, C-H absorbs around 2850–3000 cm⁻¹, converting into molecular only if the vibration induces a change. In metals, free is prominent, involving intraband transitions of conduction electrons accelerated by the of the , often assisted by ; this scales with and follows \alpha \propto \lambda^2 to \lambda^3 in the near-IR, contributing to the material's opacity. Transmission, or transmissivity T, quantifies the fraction of incident power that passes through the material unscathed. In non-scattering , energy conservation dictates that T = 1 - R - A, where R is the (fraction reflected at interfaces) and A is the (fraction absorbed internally, A = 1 - e^{-\alpha d}); for optically thin samples with low , this approximates to T \approx (1 - R)(1 - A), with R primarily determined by Fresnel reflection at the surfaces based on refractive index mismatch. Factors such as sample thickness d and surface quality thus critically influence T, as thicker samples increase A while polished surfaces minimize additional scattering losses. In semiconductors, the nature of the bandgap profoundly affects the —the wavelength threshold for significant absorption. bandgap materials, like , enable vertical electronic transitions with momentum conservation satisfied by the photon's negligible momentum, yielding a sharp absorption onset and high α (>10⁴ cm⁻¹) just above E_g. In contrast, indirect bandgap materials, such as , require involvement to conserve momentum during the transition, resulting in a weaker, more gradual with lower α (~10²–10³ cm⁻¹) and temperature-dependent behavior due to population changes.

Nonlinear Optical Properties

Second-Order Effects

Second-order nonlinear optical effects arise from the second-order susceptibility tensor \chi^{(2)}, which describes the quadratic response of the material's to the applied . The induced component is given by P_i^{(2)} = \epsilon_0 \chi^{(2)}_{ijk} E_j E_k, where \epsilon_0 is the , and \chi^{(2)}_{ijk} is a third-rank tensor with 27 elements in general, though reduces the independent components. This tensor vanishes in centrosymmetric materials due to the lack of inversion , restricting second-order effects to non-centrosymmetric belonging to the 20 point groups without a center of inversion. A prominent second-order process is (SHG), where two photons of frequency \omega interact to produce a photon at $2\omega, effectively doubling the optical frequency. Efficient SHG requires to compensate for wavevector mismatch \Delta k = k_{2\omega} - 2k_\omega, which can be achieved through in anisotropic crystals, where the and refractive indices align the phase velocities, or via quasi-phase-matching, involving periodic poling to reverse the and reset the phase every . The , or linear electro-optic effect, manifests as a change in the proportional to an applied E, expressed approximately as \Delta n = r n^3 E, with r denoting the Pockels coefficient, a material-specific tensor component. This effect enables fast of phase or polarization and is also confined to non-centrosymmetric materials, with the index deformation described by changes in the impermeability tensor \Delta (1/n^2)_{ij} = r_{ijk} E_k. Examples include SHG in quartz, where the small \chi^{(2)} (on the order of 0.1 pm/V) enables low-power frequency doubling for visible light generation, and in potassium dihydrogen phosphate (KDP) crystals, which exhibit good efficiency due to favorable phase-matching via birefringence, high damage threshold, and availability in large sizes. These processes find applications in frequency conversion for laser sources, such as generating green light from infrared Nd:YAG lasers at 1064 nm to 532 nm.

Third-Order Effects

Third-order nonlinear optical effects originate from the third-order nonlinear tensor \chi^{(3)}, which describes the response of a material to the product of three and is nonzero in all media, including centrosymmetric and isotropic ones, due to the even-rank nature of the tensor. These effects enable a variety of intensity-dependent phenomena, such as mixing and nonlinear , that are crucial for understanding light-matter interactions at high intensities. Unlike second-order processes, third-order effects do not require noncentrosymmetry, making them ubiquitous across material classes. The represents a fundamental third-order nonlinearity manifesting as an intensity-dependent , expressed as n = n_0 + n_2 I, where n_0 is the linear , I is the , and n_2 is the nonlinear coefficient proportional to the real part of \chi^{(3)}. This variation in leads to (SPM) during pulse propagation, where the instantaneous phase shift is \phi = \frac{2\pi}{\lambda} n_2 I L, with L the interaction length and \lambda the , resulting in chirped pulses and spectral broadening. The optical was first experimentally observed in liquids using intense pulses, demonstrating induced by the light field itself. Four-wave mixing (FWM) is another key third-order process driven by \chi^{(3)}, involving the interaction of three input waves at frequencies \omega_1, \omega_2, and \omega_3 to generate a fourth wave at \omega_4 = \omega_1 + \omega_2 - \omega_3. This parametric mixing transfers energy between waves and produces new frequencies, with efficiency governed by phase-matching condition \Delta k = k_1 + k_2 - k_3 - k_4 = 0, where k_i are the wave vectors; deviations from phase-matching reduce output due to wavevector mismatch. FWM was theoretically and experimentally established as a core nonlinear interaction in dielectrics, enabling applications in wavelength conversion and spectroscopy. Third-harmonic generation (THG) arises from the coherent interaction where three photons at frequency \omega produce one at $3\omega, with the nonlinear polarization P^{(3)}(3\omega) = \epsilon_0 \chi^{(3)} E^3(\omega). This process is particularly valuable in isotropic materials lacking second-order nonlinearity \chi^{(2)}, as it relies solely on \chi^{(3)} and can be phase-matched using dispersion or waveguides for enhanced conversion. Early observations confirmed THG in gases and crystals, highlighting saturation effects at high intensities that limit efficiency. Two-photon absorption (TPA) corresponds to the imaginary part of \chi^{(3)}, resulting in nonlinear absorption where the absorption coefficient becomes \alpha = \alpha_0 + \beta I, with \beta the TPA coefficient related to \operatorname{Im}(\chi^{(3)}). At the microscopic level, this equates to an excitation rate proportional to \sigma I, where \sigma is the TPA cross-section per molecule and I the intensity, enabling absorption of two lower-energy photons to reach . TPA was first demonstrated in europium-doped , showing quadratic dependence on intensity and fluorescence from the intermediate state. This effect extends linear absorption nonlinearly, with cross-sections typically on the order of $10^{-50} to $10^{-48} cm⁴ s/photon in organic dyes for near-IR excitation.

Material-Specific Properties

Dielectrics and Glasses

Dielectrics and glasses are insulating materials characterized by their wide electronic bandgaps, which enable high transparency across the visible spectrum and low optical absorption at photon energies below the bandgap. For instance, fused silica (SiO₂), a common amorphous glass, exhibits a refractive index of approximately 1.46 in the visible range and negligible absorption for wavelengths longer than about 140 nm, corresponding to a bandgap of around 9 eV. This transparency arises from the absence of free charge carriers, allowing light to propagate with minimal scattering or dissipation in these non-conductive media. Dispersion in glasses, which describes the wavelength dependence of the refractive index, is often modeled using the Sellmeier equation to account for resonant contributions from ultraviolet and infrared absorption bands. For fused silica, the Sellmeier dispersion formula is given by n^2(\lambda) = 1 + \frac{0.6961663 \lambda^2}{\lambda^2 - 0.0684043^2} + \frac{0.4079426 \lambda^2}{\lambda^2 - 0.1162414^2} + \frac{0.8974794 \lambda^2}{\lambda^2 - 9.896161^2}, where \lambda is the in micrometers and n is the ; these coefficients were derived from interferometric measurements across the 0.21–3.71 μm range. This model accurately predicts the material's normal in the visible and near-infrared, where the increases with decreasing due to closer proximity to the resonance. Birefringence, or optical , manifests in dielectrics either inherently due to or induced by external in glasses. In anisotropic crystals like (CaCO₃), the ordinary n_o is 1.658 and the extraordinary index n_e is 1.486 at the sodium D line (589 nm), resulting in a \Delta n = n_o - n_e \approx 0.172. This double-refraction enables applications such as polarizing prisms, where light splits into orthogonally polarized components following different paths. In isotropic glasses like fused silica, birefringence can be photoelastically induced by mechanical , altering the refractive index tensor and producing retardance proportional to the applied . Periodic dielectric structures, known as photonic crystals, exploit the principles of wave interference in dielectrics to engineer photonic bandgaps, ranges of frequencies where light propagation is forbidden regardless of direction. These bandgaps emerge from Bragg scattering at the periodic interfaces between high- and low-index dielectrics, analogous to electronic bandgaps in semiconductors. Seminal theoretical work demonstrated that three-dimensional photonic crystals in dielectrics could inhibit spontaneous emission and confine light on scales below the wavelength. Such structures, often fabricated from glasses or dielectric ceramics, enable applications in optical waveguides, filters, and cavities by tailoring the density of photonic states.

Metals and Semiconductors

Metals exhibit optical properties dominated by the collective behavior of free electrons, leading to high reflectivity in the . The describes this response, treating electrons as a classical gas to , which yields a permittivity of the form \epsilon(\omega) = \epsilon_\infty - \frac{\omega_p^2}{\omega(\omega + i/\tau)}, where \omega_p = \sqrt{\frac{n e^2}{\epsilon_0 m}} is the plasma frequency, n is the , e and m are the electron charge and mass, \epsilon_0 is the , \tau is the relaxation time, and \epsilon_\infty accounts for polarizability. For frequencies below \omega_p, typically in the ultraviolet for noble metals like silver and gold, the real part of the dielectric function is negative, resulting in near-total reflection with reflectivity R \approx 1 across the visible range due to the impedance mismatch with air. In contrast, semiconductors display optical properties governed by interband transitions across a bandgap E_g, with absorption beginning sharply at the absorption edge corresponding to E_g. For example, has an indirect bandgap of approximately 1.1 at , requiring assistance for momentum conservation, which leads to weaker near the edge compared to direct-gap materials like . Additionally, free-carrier in semiconductors, arising from intraband transitions of doped carriers, follows a Drude-like dependence where the \alpha \propto \lambda^2 in the , increasing with carrier concentration and due to enhanced scattering. Plasmonic effects further distinguish metallic optical responses, particularly in nanostructured forms. Surface plasmons, collective oscillations of electrons at metal-dielectric interfaces, enable strong light-matter interactions; in nanoparticles, resonance occurs around 520 nm for spheres of 20–50 nm diameter, enhancing extinction through a lineshape in the absorption and cross-sections as described by the quasi-static approximation. This resonance arises from the matching of the incident light frequency to the plasmon mode, with the peak position tunable by particle shape and surrounding medium. Doping in semiconductors modifies these properties via the Burstein-Moss shift, where high n-type doping fills the conduction band states, blocking low-energy transitions and effectively increasing the observed bandgap by an amount \Delta E_g \approx \frac{\hbar^2 (3\pi^2 n)^{2/3}}{2 m^*}, with m^* the effective mass. This shift, first observed in degenerate semiconductors like n-doped indium antimonide, enhances transparency in the near-infrared for applications in optoelectronics.

Measurement Techniques

Spectrophotometry

is a fundamental technique for quantifying the optical properties of materials by measuring the intensity of transmitted through, absorbed by, or reflected from a sample as a function of . The principle involves directing monochromatic from a source, typically covering the ultraviolet (UV), visible (Vis), and near-infrared (NIR) ranges (approximately 200 nm to 2500 nm), through the sample and detecting the resulting intensity with a . This allows for the determination of spectra, which reveal transitions, and or spectra, which indicate material or reflectivity. Common configurations include single-beam and double-beam spectrophotometers. In a single-beam setup, the light passes sequentially through the reference and sample, requiring separate measurements to compute the ratio of intensities, which can introduce errors from source fluctuations. Double-beam instruments split the light into two paths—one through the sample and one through a reference—simultaneously, enabling real-time baseline correction by subtracting or empty effects and compensating for instrumental drift. For measuring , especially in scattering samples like powders or opaque materials, an is employed; it collects light scattered in all directions by multiple internal reflections off a highly reflective , providing total data. Data analysis begins with extracting the absorption coefficient α(λ) from transmittance T, defined as the ratio of transmitted intensity to incident intensity, using the relation α(λ) = -(1/d) ln(T), where d is the sample thickness, assuming negligible reflection for thin, non-scattering samples. For more accurate results accounting for surface reflections, the formula adjusts to α(λ) = (1/d) ln[(1 - R)^2 / T], with R as reflectance. In diffuse reflection measurements of powders, the Kubelka-Munk theory models light propagation through scattering media, yielding the remission function F(R_∞) = (1 - R_∞)^2 / (2 R_∞), where R_∞ is the infinite thickness reflectance; this function is proportional to the absorption coefficient over the scattering coefficient, facilitating quantitative analysis of weakly absorbing species. A key application is bandgap determination in semiconductors, where the is analyzed via a . For indirect bandgap materials, the plot of (α hν)^{1/2} versus hν yields a linear region whose to the energy axis gives the bandgap value, providing insight into electronic structure without crystalline perfection assumptions. This method, widely used for materials like , relies on UV-Vis data near the absorption onset.

Ellipsometry and Polarimetry

Ellipsometry is a powerful optical technique that measures the change in polarization state of light upon reflection from a sample surface, providing sensitive information about thin films, interfaces, and material optical constants. It quantifies this change through two parameters: the amplitude ratio Ψ, defined as \tan \Psi = |r_p / r_s|, where r_p and r_s are the complex Fresnel reflection coefficients for p- and s-polarized light, respectively, and the phase difference Δ, given by \Delta = \arg(r_p / r_s). The Fresnel coefficient for p-polarization is expressed as r_p = \frac{n \cos \theta_i - \cos \theta_t}{n \cos \theta_i + \cos \theta_t}, where n is the refractive index, \theta_i is the angle of incidence, and \theta_t is the angle of transmission determined by Snell's law. These parameters enable the determination of refractive index, extinction coefficient, and film thickness with sub-nanometer precision, without direct contact. In applications, excels at characterizing thin-film thicknesses, such as layers on substrates, where multiple-angle measurements allow simultaneous extraction of thickness and optical constants by fitting to layered models based on the . For anisotropic materials, generalized extends the technique by incorporating direction-dependent reflection coefficients, enabling the mapping of and dichroism in structures like liquid crystals or oriented polymers. Spectroscopic variants, scanning wavelengths from to , further reveal dispersion and absorption features, making it indispensable for and metrology. Polarimetry complements ellipsometry by focusing on the rotation or modification of the polarization plane as light transmits through a material, particularly useful for probing magneto-optical and chiral properties. In the Faraday effect, a magnetic field induces a nonreciprocal rotation of the polarization plane, quantified by \theta = V B d, where V is the Verdet constant, B is the magnetic field strength, and d is the path length through the material; this is widely applied to measure magnetic properties in glasses and garnets. For chiral materials, polarimetry detects circular dichroism, the differential absorption of left- and right-circularly polarized light, which reveals molecular handedness in biomolecules and organic films. Imaging variants of these techniques, such as Mueller matrix ellipsometry, provide spatially resolved analysis by measuring the full 4x4 Mueller matrix that describes transformations of the Stokes for any input state. This approach captures , , and diattenuation across a sample, enabling high-resolution mapping of inhomogeneous thin films and nanostructures, as demonstrated in studies of black phosphorus flakes. Such methods extend traditional point measurements to microscopic scales, enhancing applications in materials .

References

  1. [1]
    None
    ### Overview of Optical Properties of Solids
  2. [2]
    None
    ### Extracted Introductory Definition and Key Concepts
  3. [3]
    Optical Properties of Materials | NIST
    Feb 26, 2021 · Such measurements include reflectance, transmittance, emittance, absorptance, and index of refraction. Many of these quantities depend upon the ...Missing: definition key aspects
  4. [4]
    [PDF] Optical Properties of Solids - UF Physics
    This book explains optical properties of solids, focusing on intrinsic properties and photoelectric emission, and is a stepping stone to advanced topics.Missing: key | Show results with:key
  5. [5]
    Nonlinear-optical properties of semiconductor composite materials
    In this paper we study the nonlinear-optical properties for the model system of CdS microcrystallites in a linear dielectric medium.
  6. [6]
    [PDF] Isotropic vs Anisotropic Minerals
    Isotropic minerals have the same chemical bonding in every direction, and consequently have the same refractive index in every direction.
  7. [7]
    Optics and Photonics (OP) | NSF - National Science Foundation
    Aug 21, 2015 · Optics and photonics are key enabling technologies that impact society in a multitude of areas, including information and communications, ...
  8. [8]
    The Science of Color - Smithsonian Libraries
    In the 1660s, English physicist and mathematician Isaac Newton began a series of experiments with sunlight and prisms. He demonstrated that clear white ...
  9. [9]
    Huygens Discovers the Wave or Pulse Theory of Light
    Huygens conceived of light as an irregular series of shock waves or pulses proceeding with very great but finite velocity through the ether.
  10. [10]
    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.
  11. [11]
    Michael Faraday's magneto-optical apparatus | Royal Institution
    His experiment was carried out in 1845 in the former servants' hall in the basement of the Royal Institution's building in 21 Albemarle Street, London.
  12. [12]
    2020: 60 Years of Lasers - Optica
    Theodore Maiman fired the first working laser. The pink ruby laser emitted mankind's first coherent light, and the world was forever changed.Missing: invention | Show results with:invention
  13. [13]
    [PDF] References: Snel's Law and Refraction Index of Refraction for Air
    The index of refraction for a medium is defined as m = c/˜c,. (1) where c is the speed of light in a vacuum, and ˜c is the speed of light in the medium.1.
  14. [14]
    [PDF] Direct method of extracting complex refractive index from routine ...
    Throughout this work, we will define the complex refractive index, en, as en = n + iκ where the real part, n, is commonly referred to simply as the ...
  15. [15]
    Anomalous Dispersion - Ocean Optics Web Book
    Oct 14, 2021 · This reversal of “colors” from that seen in normal life at visible wavelengths is called “anomalous dispersion.”
  16. [16]
    What is Cauchy dispersion module? - HORIBA
    Theoretical Model. Equation of Cauchy Transparent. The earliest dispersion formula was established by Cauchy (1836) who set up simple empirical dispersion ...
  17. [17]
    (INVITED) Methods for determining the refractive indices and thermo ...
    The Sellmeier equation, first derived in 1871 [60], is one of the most widely used representations of refractive index dispersion. Like the Cauchy equation, the ...
  18. [18]
    Principles of Birefringence | Nikon's MicroscopyU
    Uniaxial crystals have a single optical axis and comprise the largest family of common birefringent specimens, including calcite, quartz, and ordered synthetic ...
  19. [19]
    [PDF] Birefringence.pdf
    Biaxial, Uniaxial, and Isotropic Crystals. Crystals in which the three principal refractive indexes are different are termed biaxial. For crystals with ...
  20. [20]
  21. [21]
    Beer–Lambert Law - RP Photonics
    It states that the absorption coefficient is a sum of the concentration densities of all absorbing substances, each multiplied by its absorption cross-section.
  22. [22]
    Absorption Coefficient | PVEducation
    The absorption coefficient, α, in a variety of semiconductor materials ... α = 4 π k λ. where λ is the wavelength. If λ is in nm, multiply by 107 to get ...Missing: 4π | Show results with:4π
  23. [23]
    Absorption of Light - PVEducation
    When the energy of a photon is equal to or greater than the band gap of the material, the photon is absorbed by the material and excites an electron into the ...
  24. [24]
    [PDF] Infrared Spectroscopy: Theory - Organic Chemistry at CU Boulder
    Infrared radiation is absorbed by organic molecules and converted into energy of molecular vibration. In IR spectroscopy, an organic molecule is exposed to ...
  25. [25]
    Carrier Absorption - an overview | ScienceDirect Topics
    Carrier absorption is defined as the process in semiconductors where free carriers generated by linear absorption may undergo further phonon-assisted absorption ...
  26. [26]
    Transmissivity – complex transmission coefficient - RP Photonics
    The transmissivity of a surface (eg an interface between two optical media) is defined as the ratio of transmitted optical power to the incident optical power.
  27. [27]
    Direct and Indirect Band Gap Semiconductors - DoITPoMS
    In a direct band gap semiconductor, the top of the valence band and the bottom of the conduction band occur at the same value of momentum.
  28. [28]
    [PDF] The Nonlinear Optical Susceptibility - Elsevier
    Mar 1, 2014 · The quantities χ(2) and χ(3) are known as the second- and third-order non- linear optical susceptibilities, respectively. For simplicity, we ...<|control11|><|separator|>
  29. [29]
    [PDF] ICTP/ICO Winter College on Optics 1998
    The second-order susceptibility tensor must therefore vanish in all centrosymmetric materials. This is a very powerful rule because 11 of the 32 point groups ...
  30. [30]
    Phase Matching - RP Photonics
    A special technique of significant importance is quasi-phase matching, where real phase matching does not occur, but high conversion efficiencies are ...
  31. [31]
    Pockels Effect – electro-optic effect - RP Photonics
    The Pockels effect is the phenomenon that the refractive index of a medium exhibits a modification proportional to the strength of an applied electric ...
  32. [32]
    Pockels Effect - an overview | ScienceDirect Topics
    The Pockels effect is defined as the change in the refractive index of a material due to the application of a static DC electric field, which modifies the ...
  33. [33]
    Electronic structure of warm dense silicon dioxide | Phys. Rev. B
    Jun 17, 2015 · Silicon dioxide is an insulator whose band gap, in both amorphous and crystalline phases, is 8.9 eV [18, 19] . This greatly exceeds the energy ...Missing: SiO2 | Show results with:SiO2
  34. [34]
    Refractive index of Fused silica (fused quartz) - RefractiveIndex.INFO
    It has superior optical clarity, especially in the ultraviolet (UV) range, and is resistant to thermal shock, making it valuable for many high-end optical ...Missing: source | Show results with:source
  35. [35]
    2.1 The Material Silicon Dioxide - IuE
    Energy bandgap, 8.9 eV. DC resistivity, $ \approx$ 10 $ ^{17}$ $ \Omega\cdot ... These phases are often named as fused silica. Crystalline forms of SiO ...
  36. [36]
    Properties of fused silica - Heraeus Covantics
    Silicon Dioxide – Glass – Quartz – Fused Silica. Silicon Dioxide (SiO2) ... energy exceeds the network bandgap energy. The intrinsic IR edge arises due ...
  37. [37]
    Interspecimen Comparison of the Refractive Index of Fused Silica*,†
    In fused silica the influence of impurities can shift both the ultraviolet and the infrared absorption edges and this would tend to raise the refractive index, ...
  38. [38]
    Single‐scattering modeling of thin, birefringent mineral‐dust flakes ...
    Apr 11, 2009 · At λ = 589 nm wavelength, the ordinary and extraordinary refractive indices of calcite are ne = 1.486 and no = 1.658, resulting in a ...
  39. [39]
    [PDF] Photonic crystals - E. YABLONOVITCH Department of Electrical ...
    Abstract. In this paper, we review the early motivation for photonic crystal research which was derived from the need for a photonic bandgap in quantum.
  40. [40]
    [PDF] SOLID STATE PHYSICS PART II Optical Properties of Solids - MIT
    5.3 Free Carrier Absorption in Metals . ... 5.3 Free Carrier Absorption in Metals. The typical limits for metals are somewhat different than for ...
  41. [41]
    [PDF] Chapter 1 - Optical Properties of Plasmonic Materials - SPIE
    The optical properties are evaluated based on the permittivity and permeability defined by either the Drude or Lorentz model. Furthermore, the noble metals are ...
  42. [42]
    Optical Properties of Silicon - PVEducation
    The drop in absorption at the band gap (around 1100 nm) is sharper than might first appear. See also absorption coefficient. Absorption depth of silicon.png.Missing: 1.1 eV
  43. [43]
    Intrinsic shape of free carrier absorption spectra in 4H-SiC
    Jun 10, 2019 · The obtained dependences are consistent with the lambda square law predicted by the classical Drude model. The difference between the two ...Missing: proportional squared
  44. [44]
    Calculated Absorption and Scattering Properties of Gold ...
    The variation in the plasmon wavelength maximum of nanospheres, i.e., from ∼520 to 550 nm, is however too limited to be useful for in vivo applications. Gold ...
  45. [45]
    Band gap renormalization and Burstein-Moss effect in silicon
    Aug 18, 2014 · The band gap renormalization is the dominating effect up to an electron density of about ; at higher values the Burstein-Moss effect is stronger ...
  46. [46]
  47. [47]
    UV-Vis Spectroscopy: Principle, Strengths and Limitations and ...
    Dec 18, 2023 · UV-Vis spectroscopy is an analytical technique that measures the amount of discrete wavelengths of UV or visible light that are absorbed by or transmitted ...
  48. [48]
    Single Beam UV Visible Spectrophotometer vs. Double ... - Metash
    Nov 18, 2024 · Compensates for drift, offering higher stability. Baseline Correction. Manual baseline measurement required. Automated baseline correction is ...
  49. [49]
    [PDF] Integrating Sphere Theory and Applications - labsphere
    For an integrating sphere, the radiance equation must consider both multiple surface reflections and losses through the port openings needed to admit the input.
  50. [50]
    Integrating Spheres - Shimadzu
    Integrating spheres are mainly used to measure samples having scattering properties or samples such as optical lenses that refract light.
  51. [51]
    Absorbance Spectroscopy & Spectra Explained | Ossila
    The absorption coefficient, α(λ) relates to absorbance measurements through the following equation: Absorption Coefficient equation. Here, d is the distance ...
  52. [52]
    [PDF] The interpretation of diffuse reflectance spectra
    The Kubelka-Munk theory is in fact quite general, and encompasses many other two-constant theories which have been derived to suit certain select ex- perimental ...<|control11|><|separator|>
  53. [53]
    [PDF] Band Gap Determination from Diffuse Reflectance Spectra - Shimadzu
    The Tauc plot is a method that is widely used for the determination of band gap. Shown next is the procedure for determining the band gap from a diffuse.
  54. [54]
    [PDF] Band Gap Analysis through UV-Visible Spectroscopy
    Traditionally, the band gap energy is determined through use of a Tauc analysis, by which the band gap energy is related to the absorption coefficient through ...
  55. [55]
    Spectroscopic ellipsometry and polarimetry for materials and ... - NIH
    This paper discusses the fundamentals, applications, potential, limitations, and future perspectives of polarized light reflection techniques for the ...Missing: seminal | Show results with:seminal
  56. [56]
    Fundamentals and applications of spectroscopic ellipsometry - SciELO
    This paper describes the use of ellipsometry as a precise and accurate technique for characterizing substrates and overlayers.Missing: seminal | Show results with:seminal
  57. [57]
    [PDF] The Faraday Effect - Rutgers Physics
    In the visible range the refractive index of common substances, such as air, water; lead and soda glasses, etc., decreases rapidly with increasing wavelength ( ...
  58. [58]
    Mueller Matrix Ellipsometric Approach on the Imaging of Sub ...
    Jan 20, 2022 · In imaging Mueller matrix ellipsometry, an imaging system is integrated into the analyzing arm of the ellipsometer. This way, the influence of ...Introduction · Materials and Methods · Results · Discussion