Refractive index

Last updated
A ray of light being refracted through a glass slab Refraction photo.png
A ray of light being refracted through a glass slab

In optics, the refractive index (or refraction index) of an optical medium is a dimensionless number that gives the indication of the light bending ability of that medium.

Contents

Refraction of a light ray Refraction at interface.svg
Refraction of a light ray

The refractive index determines how much the path of light is bent, or refracted, when entering a material. This is described by Snell's law of refraction, n1 sin θ1 = n2 sin θ2, where θ1 and θ2 are the angle of incidence and angle of refraction, respectively, of a ray crossing the interface between two media with refractive indices n1 and n2. The refractive indices also determine the amount of light that is reflected when reaching the interface, as well as the critical angle for total internal reflection, their intensity (Fresnel equations) and Brewster's angle. [1]

The refractive index, , can be seen as the factor by which the speed and the wavelength of the radiation are reduced with respect to their vacuum values: the speed of light in a medium is v = c/n, and similarly the wavelength in that medium is λ = λ0/n, where λ0 is the wavelength of that light in vacuum. This implies that vacuum has a refractive index of 1, and assumes that the frequency (f = v/λ) of the wave is not affected by the refractive index.

The refractive index may vary with wavelength. This causes white light to split into constituent colors when refracted. This is called dispersion. This effect can be observed in prisms and rainbows, and as chromatic aberration in lenses. Light propagation in absorbing materials can be described using a complex-valued refractive index. [2] The imaginary part then handles the attenuation, while the real part accounts for refraction. For most materials the refractive index changes with wavelength by several percent across the visible spectrum. Consequently, refractive indices for materials reported using a single value for n must specify the wavelength used in the measurement.

The concept of refractive index applies across the full electromagnetic spectrum, from X-rays to radio waves. It can also be applied to wave phenomena such as sound. In this case, the speed of sound is used instead of that of light, and a reference medium other than vacuum must be chosen. [3]

For lenses (such as eye glasses), a lens made from a high refractive index material will be thinner, and hence lighter, than a conventional lens with a lower refractive index. Such lenses are generally more expensive to manufacture than conventional ones.

Definition

The relative refractive index of an optical medium 2 with respect to another reference medium 1 (n21) is given by the ratio of speed of light in medium 1 to that in medium 2. This can be expressed as follows:

If the reference medium 1 is vacuum, then the refractive index of medium 2 is considered with respect to vacuum. It is simply represented as n2 and is called the absolute refractive index of medium 2.

The absolute refractive indexn of an optical medium is defined as the ratio of the speed of light in vacuum, c = 299792458 m/s, and the phase velocity v of light in the medium,

Since c is constant, n is inversely proportional to v:

The phase velocity is the speed at which the crests or the phase of the wave moves, which may be different from the group velocity, the speed at which the pulse of light or the envelope of the wave moves. [1] Historically air at a standardized pressure and temperature has been common as a reference medium.

History

Thomas Young coined the term index of refraction in 1807. Thomas Young (scientist).jpg
Thomas Young coined the term index of refraction in 1807.

Thomas Young was presumably the person who first used, and invented, the name "index of refraction", in 1807. [4] At the same time he changed this value of refractive power into a single number, instead of the traditional ratio of two numbers. The ratio had the disadvantage of different appearances. Newton, who called it the "proportion of the sines of incidence and refraction", wrote it as a ratio of two numbers, like "529 to 396" (or "nearly 4 to 3"; for water). [5] Hauksbee, who called it the "ratio of refraction", wrote it as a ratio with a fixed numerator, like "10000 to 7451.9" (for urine). [6] Hutton wrote it as a ratio with a fixed denominator, like 1.3358 to 1 (water). [7]

Young did not use a symbol for the index of refraction, in 1807. In the later years, others started using different symbols: n, m, and µ. [8] [9] [10] The symbol n gradually prevailed.

Typical values

Diamonds have a very high refractive index of 2.417. Brillanten.jpg
Diamonds have a very high refractive index of 2.417.

Refractive index also varies with wavelength of the light as given by Cauchy's equation:

The most general form of Cauchy's equation is

where n is the refractive index, λ is the wavelength, and A, B, C, etc., are coefficients that can be determined for a material by fitting the equation to measured refractive indices at known wavelengths. The coefficients are usually quoted for λ as the vacuum wavelength in micrometres.

Usually, it is sufficient to use a two-term form of the equation:

where the coefficients A and B are determined specifically for this form of the equation.

Selected refractive indices at λ = 589 nm. For references, see the extended List of refractive indices.
Materialn
Vacuum 1
Gases at 0 °C and 1 atm
Air 1.000293
Helium 1.000036
Hydrogen 1.000132
Carbon dioxide 1.00045
Liquids at 20 °C
Water 1.333
Ethanol 1.36
Olive oil 1.47
Solids
Ice 1.31
Fused silica (quartz)1.46 [11]
PMMA (acrylic, plexiglas, lucite, perspex)1.49
Window glass 1.52 [12]
Polycarbonate (Lexan™)1.58 [13]
Flint glass (typical)1.69
Sapphire 1.77 [14]
Cubic zirconia 2.15
Diamond 2.417
Moissanite 2.65

For visible light most transparent media have refractive indices between 1 and 2. A few examples are given in the adjacent table. These values are measured at the yellow doublet D-line of sodium, with a wavelength of 589 nanometers, as is conventionally done. [15] Gases at atmospheric pressure have refractive indices close to 1 because of their low density. Almost all solids and liquids have refractive indices above 1.3, with aerogel as the clear exception. Aerogel is a very low density solid that can be produced with refractive index in the range from 1.002 to 1.265. [16] Moissanite lies at the other end of the range with a refractive index as high as 2.65. Most plastics have refractive indices in the range from 1.3 to 1.7, but some high-refractive-index polymers can have values as high as 1.76. [17]

For infrared light refractive indices can be considerably higher. Germanium is transparent in the wavelength region from 2 to 14 μm and has a refractive index of about 4. [18] A type of new materials termed "topological insulators", was recently found which have high refractive index of up to 6 in the near to mid infrared frequency range. Moreover, topological insulators are transparent when they have nanoscale thickness. These properties are potentially important for applications in infrared optics. [19]

Refractive index below unity

According to the theory of relativity, no information can travel faster than the speed of light in vacuum, but this does not mean that the refractive index cannot be less than 1. The refractive index measures the phase velocity of light, which does not carry information. [20] [lower-alpha 1] The phase velocity is the speed at which the crests of the wave move and can be faster than the speed of light in vacuum, and thereby give a refractive index below 1. This can occur close to resonance frequencies, for absorbing media, in plasmas, and for X-rays. In the X-ray regime the refractive indices are lower than but very close to 1 (exceptions close to some resonance frequencies). [21] As an example, water has a refractive index of 0.99999974 =1 − 2.6×10−7 for X-ray radiation at a photon energy of 30  keV (0.04 nm wavelength). [21]

An example of a plasma with an index of refraction less than unity is Earth's ionosphere. Since the refractive index of the ionosphere (a plasma), is less than unity, electromagnetic waves propagating through the plasma are bent "away from the normal" (see Geometric optics) allowing the radio wave to be refracted back toward earth, thus enabling long-distance radio communications. See also Radio Propagation and Skywave. [22]

Negative refractive index

A split-ring resonator array arranged to produce a negative index of refraction for microwaves Split-ring resonator array 10K sq nm.jpg
A split-ring resonator array arranged to produce a negative index of refraction for microwaves

Recent research has also demonstrated the "existence" of materials with a negative refractive index, which can occur if permittivity and permeability have simultaneous negative values. [23] This can be achieved with periodically constructed metamaterials. The resulting negative refraction (i.e., a reversal of Snell's law) offers the possibility of the superlens and other new phenomena to be actively developed by means of metamaterials. [24] [25]

Microscopic explanation

In optical mineralogy, thin sections are used to study rocks. The method is based on the distinct refractive indices of different minerals. Thin section scan crossed polarizers Siilinjarvi R636-105.90.jpg
In optical mineralogy, thin sections are used to study rocks. The method is based on the distinct refractive indices of different minerals.

At the atomic scale, an electromagnetic wave's phase velocity is slowed in a material because the electric field creates a disturbance in the charges of each atom (primarily the electrons) proportional to the electric susceptibility of the medium. (Similarly, the magnetic field creates a disturbance proportional to the magnetic susceptibility.) As the electromagnetic fields oscillate in the wave, the charges in the material will be "shaken" back and forth at the same frequency. [1] :67 The charges thus radiate their own electromagnetic wave that is at the same frequency, but usually with a phase delay, as the charges may move out of phase with the force driving them (see sinusoidally driven harmonic oscillator). The light wave traveling in the medium is the macroscopic superposition (sum) of all such contributions in the material: the original wave plus the waves radiated by all the moving charges. This wave is typically a wave with the same frequency but shorter wavelength than the original, leading to a slowing of the wave's phase velocity. Most of the radiation from oscillating material charges will modify the incoming wave, changing its velocity. However, some net energy will be radiated in other directions or even at other frequencies (see scattering).

Depending on the relative phase of the original driving wave and the waves radiated by the charge motion, there are several possibilities:

For most materials at visible-light frequencies, the phase is somewhere between 90° and 180°, corresponding to a combination of both refraction and absorption.

Dispersion

Light of different colors has slightly different refractive indices in water and therefore shows up at different positions in the rainbow. WhereRainbowRises.jpg
Light of different colors has slightly different refractive indices in water and therefore shows up at different positions in the rainbow.
In a triangular prism, dispersion causes different colors to refract at different angles, splitting white light into a rainbow of colors. The blue color is more deviated (refracted) than the red color because the refractive index of blue is higher than that of red. Prism-rainbow.svg
In a triangular prism, dispersion causes different colors to refract at different angles, splitting white light into a rainbow of colors. The blue color is more deviated (refracted) than the red color because the refractive index of blue is higher than that of red.
The variation of refractive index with wavelength for various glasses. The shaded zone indicates the range of visible light. Mplwp dispersion curves.svg
The variation of refractive index with wavelength for various glasses. The shaded zone indicates the range of visible light.

The refractive index of materials varies with the wavelength (and frequency) of light. [27] This is called dispersion and causes prisms and rainbows to divide white light into its constituent spectral colors. [28] As the refractive index varies with wavelength, so will the refraction angle as light goes from one material to another. Dispersion also causes the focal length of lenses to be wavelength dependent. This is a type of chromatic aberration, which often needs to be corrected for in imaging systems. In regions of the spectrum where the material does not absorb light, the refractive index tends to decrease with increasing wavelength, and thus increase with frequency. This is called "normal dispersion", in contrast to "anomalous dispersion", where the refractive index increases with wavelength. [27] For visible light normal dispersion means that the refractive index is higher for blue light than for red.

For optics in the visual range, the amount of dispersion of a lens material is often quantified by the Abbe number: [28]

For a more accurate description of the wavelength dependence of the refractive index, the Sellmeier equation can be used. [29] It is an empirical formula that works well in describing dispersion. Sellmeier coefficients are often quoted instead of the refractive index in tables.

Principal refractive index wavelength ambiguity

Because of dispersion, it is usually important to specify the vacuum wavelength of light for which a refractive index is measured. Typically, measurements are done at various well-defined spectral emission lines.

Manufacturers of optical glass in general define principal index of refraction at yellow spectral line of helium (587.56 nm) and alternatively at a green spectral line of mercury (546.07 nm), called d and e lines respectively. Abbe number is defined for both and denoted Vd and Ve. The spectral data provided by glass manufacturers is also often more precise for these two wavelengths. [30] [31] [32] [33]

Both, d and e spectral lines are singlets and thus are suitable to perform a very precise measurements, such as spectral goniometric method. [34] [35]

In practical applications, measurements of refractive index are performed on various refractometers, such as Abbe refractometer. Measurement accuracy of such typical commercial devices is in the order of 0.0002. [36] [37] Refractometers usually measure refractive index nD, defined for sodium doublet D (589.29 nm), which is actually a midpoint between two adjacent yellow spectral lines of sodium. Yellow spectral lines of helium (d) and sodium (D) are 1.73 nm apart, which can be considered negligible for typical refractometers, but can cause confusion and lead to errors if accuracy is critical.

All three typical principle refractive indices definitions can be found depending on application and region, [38] so a proper subscript should be used to avoid ambiguity.

Complex refractive index

When light passes through a medium, some part of it will always be absorbed. This can be conveniently taken into account by defining a complex refractive index,

Here, the real part n is the refractive index and indicates the phase velocity, while the imaginary part κ is called the optical extinction coefficient or absorption coefficient —although κ can also refer to the mass attenuation coefficient [39] :3—and indicates the amount of attenuation when the electromagnetic wave propagates through the material. [1] :128

That κ corresponds to absorption can be seen by inserting this refractive index into the expression for electric field of a plane electromagnetic wave traveling in the x-direction. This can be done by relating the complex wave number k to the complex refractive index n through k = 2πn/λ0, with λ0 being the vacuum wavelength; this can be inserted into the plane wave expression for a wave travelling in the x-direction as:

Here we see that κ gives an exponential decay, as expected from the Beer–Lambert law. Since intensity is proportional to the square of the electric field, intensity will depend on the depth into the material as

and thus the absorption coefficient is α = 4πκ/λ0, [1] :128 and the penetration depth (the distance after which the intensity is reduced by a factor of 1/e) is δp = 1/α = λ0/4πκ.

Both n and κ are dependent on the frequency. In most circumstances κ > 0 (light is absorbed) or κ = 0 (light travels forever without loss). In special situations, especially in the gain medium of lasers, it is also possible that κ < 0, corresponding to an amplification of the light.

An alternative convention uses n = n + instead of n = n, but where κ > 0 still corresponds to loss. Therefore, these two conventions are inconsistent and should not be confused. The difference is related to defining sinusoidal time dependence as Re[exp(−iωt)] versus Re[exp(+iωt)]. See Mathematical descriptions of opacity.

Dielectric loss and non-zero DC conductivity in materials cause absorption. Good dielectric materials such as glass have extremely low DC conductivity, and at low frequencies the dielectric loss is also negligible, resulting in almost no absorption. However, at higher frequencies (such as visible light), dielectric loss may increase absorption significantly, reducing the material's transparency to these frequencies.

The real n, and imaginary κ, parts of the complex refractive index are related through the Kramers–Kronig relations. In 1986, A.R. Forouhi and I. Bloomer deduced an equation describing κ as a function of photon energy, E, applicable to amorphous materials. Forouhi and Bloomer then applied the Kramers–Kronig relation to derive the corresponding equation for n as a function of E. The same formalism was applied to crystalline materials by Forouhi and Bloomer in 1988.

The refractive index and extinction coefficient, n and κ, are typically measured from quantities that depend on them, such as reflectance, R, or transmittance, T, or ellipsometric parameters, ψ and δ. The determination of n and κ from such measured quantities will involve developing a theoretical expression for R or T, or ψ and δ in terms of a valid physical model for n and κ. By fitting the theoretical model to the measured R or T, or ψ and δ using regression analysis, n and κ can be deduced.

X-ray and extreme UV

For X-ray and extreme ultraviolet radiation the complex refractive index deviates only slightly from unity and usually has a real part smaller than 1. It is therefore normally written as n = 1 − δ + (or n = 1 − δ with the alternative convention mentioned above). [2] Far above the atomic resonance frequency delta can be given by

where r0 is the classical electron radius, λ is the X-ray wavelength, and ne is the electron density. One may assume the electron density is simply the number of electrons per atom Z multiplied by the atomic density, but more accurate calculation of the refractive index requires replacing Z with the complex atomic form factor . It follows that

with δ and β typically of the order of 10−5 and 10−6.

Relations to other quantities

Optical path length

The colors of a soap bubble are determined by the optical path length through the thin soap film in a phenomenon called thin-film interference. Soap bubble sky.jpg
The colors of a soap bubble are determined by the optical path length through the thin soap film in a phenomenon called thin-film interference.

Optical path length (OPL) is the product of the geometric length d of the path light follows through a system, and the index of refraction of the medium through which it propagates, [40] This is an important concept in optics because it determines the phase of the light and governs interference and diffraction of light as it propagates. According to Fermat's principle, light rays can be characterized as those curves that optimize the optical path length. [1] :68–69

Refraction

Refraction of light at the interface between two media of different refractive indices, with n2 > n1. Since the phase velocity is lower in the second medium (v2 < v1), the angle of refraction th2 is less than the angle of incidence th1; that is, the ray in the higher-index medium is closer to the normal. Snells law.svg
Refraction of light at the interface between two media of different refractive indices, with n2 > n1. Since the phase velocity is lower in the second medium (v2 < v1), the angle of refraction θ2 is less than the angle of incidence θ1; that is, the ray in the higher-index medium is closer to the normal.

When light moves from one medium to another, it changes direction, i.e. it is refracted. If it moves from a medium with refractive index n1 to one with refractive index n2, with an incidence angle to the surface normal of θ1, the refraction angle θ2 can be calculated from Snell's law: [41]

When light enters a material with higher refractive index, the angle of refraction will be smaller than the angle of incidence and the light will be refracted towards the normal of the surface. The higher the refractive index, the closer to the normal direction the light will travel. When passing into a medium with lower refractive index, the light will instead be refracted away from the normal, towards the surface.

Total internal reflection

Total internal reflection can be seen at the air-water boundary. Total internal reflection of Chelonia mydas.jpg
Total internal reflection can be seen at the air-water boundary.

If there is no angle θ2 fulfilling Snell's law, i.e.,

the light cannot be transmitted and will instead undergo total internal reflection. [42] :49–50 This occurs only when going to a less optically dense material, i.e., one with lower refractive index. To get total internal reflection the angles of incidence θ1 must be larger than the critical angle [43]

Reflectivity

Apart from the transmitted light there is also a reflected part. The reflection angle is equal to the incidence angle, and the amount of light that is reflected is determined by the reflectivity of the surface. The reflectivity can be calculated from the refractive index and the incidence angle with the Fresnel equations, which for normal incidence reduces to [42] :44

For common glass in air, n1 = 1 and n2 = 1.5, and thus about 4% of the incident power is reflected. [44] At other incidence angles the reflectivity will also depend on the polarization of the incoming light. At a certain angle called Brewster's angle, p-polarized light (light with the electric field in the plane of incidence) will be totally transmitted. Brewster's angle can be calculated from the two refractive indices of the interface as [1] :245

Lenses

The power of a magnifying glass is determined by the shape and refractive index of the lens. Lupa.na.encyklopedii.jpg
The power of a magnifying glass is determined by the shape and refractive index of the lens.

The focal length of a lens is determined by its refractive index n and the radii of curvature R1 and R2 of its surfaces. The power of a thin lens in air is given by the simplified version of the Lensmaker's formula: [45]

where f is the focal length of the lens.

Microscope resolution

The resolution of a good optical microscope is mainly determined by the numerical aperture (ANum) of its objective lens. The numerical aperture in turn is determined by the refractive index n of the medium filling the space between the sample and the lens and the half collection angle of light θ according to Carlsson (2007): [46] :6

For this reason oil immersion is commonly used to obtain high resolution in microscopy. In this technique the objective is dipped into a drop of high refractive index immersion oil on the sample under study. [46] :14

Relative permittivity and permeability

The refractive index of electromagnetic radiation equals

where εr is the material's relative permittivity, and μr is its relative permeability. [47] :229 The refractive index is used for optics in Fresnel equations and Snell's law; while the relative permittivity and permeability are used in Maxwell's equations and electronics. Most naturally occurring materials are non-magnetic at optical frequencies, that is μr is very close to 1, therefore n is approximately εr. [48] In this particular case, the complex relative permittivity εr, with real and imaginary parts εr and ɛ̃r, and the complex refractive index n, with real and imaginary parts n and κ (the latter called the "extinction coefficient"), follow the relation

and their components are related by: [49]

and:

where is the complex modulus.

Wave impedance

The wave impedance of a plane electromagnetic wave in a non-conductive medium is given by

where Z0 is the vacuum wave impedance, μ and ε are the absolute permeability and permittivity of the medium, εr is the material's relative permittivity, and μr is its relative permeability.

In non-magnetic media (that is, in materials with μr = 1), and

Thus refractive index in a non-magnetic media is the ratio of the vacuum wave impedance to the wave impedance of the medium.

The reflectivity R0 between two media can thus be expressed both by the wave impedances and the refractive indices as

Density

The relation between the refractive index and the density of silicate and borosilicate glasses Density-nd.GIF
The relation between the refractive index and the density of silicate and borosilicate glasses

In general, the refractive index of a glass increases with its density. However, there does not exist an overall linear relationship between the refractive index and the density for all silicate and borosilicate glasses. A relatively high refractive index and low density can be obtained with glasses containing light metal oxides such as Li2O and MgO, while the opposite trend is observed with glasses containing PbO and BaO as seen in the diagram at the right.

Many oils (such as olive oil) and ethanol are examples of liquids that are more refractive, but less dense, than water, contrary to the general correlation between density and refractive index.

For air, n - 1 is proportional to the density of the gas as long as the chemical composition does not change. [51] This means that it is also proportional to the pressure and inversely proportional to the temperature for ideal gases.

Group index

Sometimes, a "group velocity refractive index", usually called the group index is defined:[ citation needed ]

where vg is the group velocity. This value should not be confused with n, which is always defined with respect to the phase velocity. When the dispersion is small, the group velocity can be linked to the phase velocity by the relation [42] :22

where λ is the wavelength in the medium. In this case the group index can thus be written in terms of the wavelength dependence of the refractive index as

When the refractive index of a medium is known as a function of the vacuum wavelength (instead of the wavelength in the medium), the corresponding expressions for the group velocity and index are (for all values of dispersion) [52]

where λ0 is the wavelength in vacuum.

Velocity, momentum, and polarizability

As shown in the Fizeau experiment, when light is transmitted through a moving medium, its speed relative to an observer traveling with speed v in the same direction as the light is:

The momentum of photons in a medium of refractive index n is a complex and controversial issue with two different values having different physical interpretations. [53]

The refractive index of a substance can be related to its polarizability with the Lorentz–Lorenz equation or to the molar refractivities of its constituents by the Gladstone–Dale relation.

Refractivity

In atmospheric applications, refractivity is defined as N = n – 1, often rescaled as either [54] N = 106 (n – 1) [55] [56] or N = 108 (n – 1); [57] the multiplication factors are used because the refractive index for air, n deviates from unity by at most a few parts per ten thousand.

Molar refractivity , on the other hand, is a measure of the total polarizability of a mole of a substance and can be calculated from the refractive index as

where ρ is the density, and M is the molar mass. [42] :93

Nonscalar, nonlinear, or nonhomogeneous refraction

So far, we have assumed that refraction is given by linear equations involving a spatially constant, scalar refractive index. These assumptions can break down in different ways, to be described in the following subsections.

Birefringence

A calcite crystal laid upon a paper with some letters showing double refraction Calcite.jpg
A calcite crystal laid upon a paper with some letters showing double refraction
Birefringent materials can give rise to colors when placed between crossed polarizers. This is the basis for photoelasticity. Plastic Protractor Polarized 05375.jpg
Birefringent materials can give rise to colors when placed between crossed polarizers. This is the basis for photoelasticity.

In some materials, the refractive index depends on the polarization and propagation direction of the light. [58] This is called birefringence or optical anisotropy.

In the simplest form, uniaxial birefringence, there is only one special direction in the material. This axis is known as the optical axis of the material. [1] :230 Light with linear polarization perpendicular to this axis will experience an ordinary refractive index no while light polarized in parallel will experience an extraordinary refractive index ne. [1] :236 The birefringence of the material is the difference between these indices of refraction, Δn = neno. [1] :237 Light propagating in the direction of the optical axis will not be affected by the birefringence since the refractive index will be no independent of polarization. For other propagation directions the light will split into two linearly polarized beams. For light traveling perpendicularly to the optical axis the beams will have the same direction. [1] :233 This can be used to change the polarization direction of linearly polarized light or to convert between linear, circular, and elliptical polarizations with waveplates. [1] :237

Many crystals are naturally birefringent, but isotropic materials such as plastics and glass can also often be made birefringent by introducing a preferred direction through, e.g., an external force or electric field. This effect is called photoelasticity, and can be used to reveal stresses in structures. The birefringent material is placed between crossed polarizers. A change in birefringence alters the polarization and thereby the fraction of light that is transmitted through the second polarizer.

In the more general case of trirefringent materials described by the field of crystal optics, the dielectric constant is a rank-2 tensor (a 3 by 3 matrix). In this case the propagation of light cannot simply be described by refractive indices except for polarizations along principal axes.

Nonlinearity

The strong electric field of high intensity light (such as the output of a laser) may cause a medium's refractive index to vary as the light passes through it, giving rise to nonlinear optics. [1] :502 If the index varies quadratically with the field (linearly with the intensity), it is called the optical Kerr effect and causes phenomena such as self-focusing and self-phase modulation. [1] :264 If the index varies linearly with the field (a nontrivial linear coefficient is only possible in materials that do not possess inversion symmetry), it is known as the Pockels effect. [1] :265

Inhomogeneity

A gradient-index lens with a parabolic variation of refractive index (n) with radial distance (x). The lens focuses light in the same way as a conventional lens. Grin-lens.png
A gradient-index lens with a parabolic variation of refractive index (n) with radial distance (x). The lens focuses light in the same way as a conventional lens.

If the refractive index of a medium is not constant but varies gradually with the position, the material is known as a gradient-index (GRIN) medium and is described by gradient index optics. [1] :273 Light traveling through such a medium can be bent or focused, and this effect can be exploited to produce lenses, some optical fibers, and other devices. Introducing GRIN elements in the design of an optical system can greatly simplify the system, reducing the number of elements by as much as a third while maintaining overall performance. [1] :276 The crystalline lens of the human eye is an example of a GRIN lens with a refractive index varying from about 1.406 in the inner core to approximately 1.386 at the less dense cortex. [1] :203 Some common mirages are caused by a spatially varying refractive index of air.

Refractive index measurement

Homogeneous media

The principle of many refractometers Pulfrich refraktometer en.png
The principle of many refractometers

The refractive index of liquids or solids can be measured with refractometers. They typically measure some angle of refraction or the critical angle for total internal reflection. The first laboratory refractometers sold commercially were developed by Ernst Abbe in the late 19th century. [59] The same principles are still used today. In this instrument, a thin layer of the liquid to be measured is placed between two prisms. Light is shone through the liquid at incidence angles all the way up to 90°, i.e., light rays parallel to the surface. The second prism should have an index of refraction higher than that of the liquid, so that light only enters the prism at angles smaller than the critical angle for total reflection. This angle can then be measured either by looking through a telescope,[ clarification needed ] or with a digital photodetector placed in the focal plane of a lens. The refractive index n of the liquid can then be calculated from the maximum transmission angle θ as n = nG sin θ, where nG is the refractive index of the prism. [60]

A handheld refractometer used to measure the sugar content of fruits Refractometer.jpg
A handheld refractometer used to measure the sugar content of fruits

This type of device is commonly used in chemical laboratories for identification of substances and for quality control. Handheld variants are used in agriculture by, e.g., wine makers to determine sugar content in grape juice, and inline process refractometers are used in, e.g., chemical and pharmaceutical industry for process control.

In gemology, a different type of refractometer is used to measure the index of refraction and birefringence of gemstones. The gem is placed on a high refractive index prism and illuminated from below. A high refractive index contact liquid is used to achieve optical contact between the gem and the prism. At small incidence angles most of the light will be transmitted into the gem, but at high angles total internal reflection will occur in the prism. The critical angle is normally measured by looking through a telescope. [61]

Refractive index variations

A differential interference contrast microscopy image of budding yeast cells S cerevisiae under DIC microscopy.jpg
A differential interference contrast microscopy image of budding yeast cells

Unstained biological structures appear mostly transparent under bright-field microscopy as most cellular structures do not attenuate appreciable quantities of light. Nevertheless, the variation in the materials that constitute these structures also corresponds to a variation in the refractive index. The following techniques convert such variation into measurable amplitude differences:

To measure the spatial variation of the refractive index in a sample phase-contrast imaging methods are used. These methods measure the variations in phase of the light wave exiting the sample. The phase is proportional to the optical path length the light ray has traversed, and thus gives a measure of the integral of the refractive index along the ray path. The phase cannot be measured directly at optical or higher frequencies, and therefore needs to be converted into intensity by interference with a reference beam. In the visual spectrum this is done using Zernike phase-contrast microscopy, differential interference contrast microscopy (DIC), or interferometry.

Zernike phase-contrast microscopy introduces a phase shift to the low spatial frequency components of the image with a phase-shifting annulus in the Fourier plane of the sample, so that high-spatial-frequency parts of the image can interfere with the low-frequency reference beam. In DIC the illumination is split up into two beams that are given different polarizations, are phase shifted differently, and are shifted transversely with slightly different amounts. After the specimen, the two parts are made to interfere, giving an image of the derivative of the optical path length in the direction of the difference in the transverse shift. [46] In interferometry the illumination is split up into two beams by a partially reflective mirror. One of the beams is let through the sample before they are combined to interfere and give a direct image of the phase shifts. If the optical path length variations are more than a wavelength the image will contain fringes.

There exist several phase-contrast X-ray imaging techniques to determine 2D or 3D spatial distribution of refractive index of samples in the X-ray regime. [62]

Applications

The refractive index is an important property of the components of any optical instrument. It determines the focusing power of lenses, the dispersive power of prisms, the reflectivity of lens coatings, and the light-guiding nature of optical fiber. Since the refractive index is a fundamental physical property of a substance, it is often used to identify a particular substance, confirm its purity, or measure its concentration. The refractive index is used to measure solids, liquids, and gases. Most commonly it is used to measure the concentration of a solute in an aqueous solution. It can also be used as a useful tool to differentiate between different types of gemstone, due to the unique chatoyance each individual stone displays. A refractometer is the instrument used to measure the refractive index. For a solution of sugar, the refractive index can be used to determine the sugar content (see Brix).

See also

Footnotes

  1. One consequence of the real part of n being less than unity is that it implies that the phase velocity inside the material, c/n, is larger than the velocity of light, c. This does not, however, violate the law of relativity, which requires that only signals carrying information do not travel faster than c. Such signals move with the group velocity, not with the phase velocity, and it can be shown that the group velocity is in fact less than c. [20]

Related Research Articles

The Beer-Lambert law is commonly applied to chemical analysis measurements to determine the concentration of chemical species that absorb light. It is often referred to as Beer's law. In physics, the Bouguer–Lambert law is an empirical law which relates the extinction or attenuation of light to the properties of the material through which the light is travelling. It had its first use in astronomical extinction. The fundamental law of extinction is sometimes called the Beer-Bouguer-Lambert law or the Bouguer-Beer-Lambert law or merely the extinction law. The extinction law is also used in understanding attenuation in physical optics, for photons, neutrons, or rarefied gases. In mathematical physics, this law arises as a solution of the BGK equation.

In physics, the cross section is a measure of the probability that a specific process will take place in a collision of two particles. For example, the Rutherford cross-section is a measure of probability that an alpha particle will be deflected by a given angle during an interaction with an atomic nucleus. Cross section is typically denoted σ (sigma) and is expressed in units of area, more specifically in barns. In a way, it can be thought of as the size of the object that the excitation must hit in order for the process to occur, but more exactly, it is a parameter of a stochastic process.

<span class="mw-page-title-main">Fresnel equations</span> Equations of light transmission and reflection

The Fresnel equations describe the reflection and transmission of light when incident on an interface between different optical media. They were deduced by French engineer and physicist Augustin-Jean Fresnel who was the first to understand that light is a transverse wave, when no one realized that the waves were electric and magnetic fields. For the first time, polarization could be understood quantitatively, as Fresnel's equations correctly predicted the differing behaviour of waves of the s and p polarizations incident upon a material interface.

<span class="mw-page-title-main">Nonlinear optics</span> Branch of physics

Nonlinear optics (NLO) is the branch of optics that describes the behaviour of light in nonlinear media, that is, media in which the polarization density P responds non-linearly to the electric field E of the light. The non-linearity is typically observed only at very high light intensities (when the electric field of the light is >108 V/m and thus comparable to the atomic electric field of ~1011 V/m) such as those provided by lasers. Above the Schwinger limit, the vacuum itself is expected to become nonlinear. In nonlinear optics, the superposition principle no longer holds.

<span class="mw-page-title-main">Refraction</span> Physical phenomenon relating to the direction of waves

In physics, refraction is the redirection of a wave as it passes from one medium to another. The redirection can be caused by the wave's change in speed or by a change in the medium. Refraction of light is the most commonly observed phenomenon, but other waves such as sound waves and water waves also experience refraction. How much a wave is refracted is determined by the change in wave speed and the initial direction of wave propagation relative to the direction of change in speed.

<span class="mw-page-title-main">Rayleigh scattering</span> Light scattering by small particles

Rayleigh scattering, named after the 19th-century British physicist Lord Rayleigh, is the predominantly elastic scattering of light, or other electromagnetic radiation, by particles with a size much smaller than the wavelength of the radiation. For light frequencies well below the resonance frequency of the scattering medium, the amount of scattering is inversely proportional to the fourth power of the wavelength, e.g., a blue color is scattered much more than a red color as light propagates through air.

<span class="mw-page-title-main">Wavelength</span> Distance over which a waves shape repeats

In physics and mathematics, wavelength or spatial period of a wave or periodic function is the distance over which the wave's shape repeats. In other words, it is the distance between consecutive corresponding points of the same phase on the wave, such as two adjacent crests, troughs, or zero crossings. Wavelength is a characteristic of both traveling waves and standing waves, as well as other spatial wave patterns. The inverse of the wavelength is called the spatial frequency. Wavelength is commonly designated by the Greek letter lambda (λ). The term "wavelength" is also sometimes applied to modulated waves, and to the sinusoidal envelopes of modulated waves or waves formed by interference of several sinusoids.

<span class="mw-page-title-main">Gaussian beam</span> Monochrome light beam whose amplitude envelope is a Gaussian function

In optics, a Gaussian beam is an idealized beam of electromagnetic radiation whose amplitude envelope in the transverse plane is given by a Gaussian function; this also implies a Gaussian intensity (irradiance) profile. This fundamental (or TEM00) transverse Gaussian mode describes the intended output of many lasers, as such a beam diverges less and can be focused better than any other. When a Gaussian beam is refocused by an ideal lens, a new Gaussian beam is produced. The electric and magnetic field amplitude profiles along a circular Gaussian beam of a given wavelength and polarization are determined by two parameters: the waistw0, which is a measure of the width of the beam at its narrowest point, and the position z relative to the waist.

<span class="mw-page-title-main">Snell's law</span> Formula for refraction angles

Snell's law is a formula used to describe the relationship between the angles of incidence and refraction, when referring to light or other waves passing through a boundary between two different isotropic media, such as water, glass, or air. In optics, the law is used in ray tracing to compute the angles of incidence or refraction, and in experimental optics to find the refractive index of a material. The law is also satisfied in meta-materials, which allow light to be bent "backward" at a negative angle of refraction with a negative refractive index.

<span class="mw-page-title-main">Relative permittivity</span> Measure of the electric polarizability of a dielectric, compared with that of a vacuum

The relative permittivity is the permittivity of a material expressed as a ratio with the electric permittivity of a vacuum. A dielectric is an insulating material, and the dielectric constant of an insulator measures the ability of the insulator to store electric energy in an electrical field.

<span class="mw-page-title-main">Permittivity</span> Measure of the electric polarizability of a dielectric

In electromagnetism, the absolute permittivity, often simply called permittivity and denoted by the Greek letter ε (epsilon), is a measure of the electric polarizability of a dielectric. A material with high permittivity polarizes more in response to an applied electric field than a material with low permittivity, thereby storing more energy in the material. In electrostatics, the permittivity plays an important role in determining the capacitance of a capacitor.

<span class="mw-page-title-main">Wavenumber</span> Spatial frequency of a wave

In the physical sciences, the wavenumber, also known as repetency, is the spatial frequency of a wave, measured in cycles per unit distance or radians per unit distance. It is analogous to temporal frequency, which is defined as the number of wave cycles per unit time or radians per unit time.

<span class="mw-page-title-main">Dispersion (optics)</span> Dependence of phase velocity on frequency

In optics and in wave propagation in general, dispersion is the phenomenon in which the phase velocity of a wave depends on its frequency; sometimes the term chromatic dispersion is used for specificity to optics in particular. A medium having this common property may be termed a dispersive medium.

<span class="mw-page-title-main">Transmittance</span> Effectiveness of a material in transmitting radiant energy

In optical physics, transmittance of the surface of a material is its effectiveness in transmitting radiant energy. It is the fraction of incident electromagnetic power that is transmitted through a sample, in contrast to the transmission coefficient, which is the ratio of the transmitted to incident electric field.

The Kerr effect, also called the quadratic electro-optic (QEO) effect, is a change in the refractive index of a material in response to an applied electric field. The Kerr effect is distinct from the Pockels effect in that the induced index change is directly proportional to the square of the electric field instead of varying linearly with it. All materials show a Kerr effect, but certain liquids display it more strongly than others. The Kerr effect was discovered in 1875 by Scottish physicist John Kerr.

Geometrical optics, or ray optics, is a model of optics that describes light propagation in terms of rays. The ray in geometrical optics is an abstraction useful for approximating the paths along which light propagates under certain circumstances.

<span class="mw-page-title-main">Metamaterial</span> Materials engineered to have properties that have not yet been found in nature

A metamaterial is any material engineered to have a property that is rarely observed in naturally occurring materials. They are made from assemblies of multiple elements fashioned from composite materials such as metals and plastics. These materials are usually arranged in repeating patterns, at scales that are smaller than the wavelengths of the phenomena they influence. Metamaterials derive their properties not from the properties of the base materials, but from their newly designed structures. Their precise shape, geometry, size, orientation and arrangement gives them their smart properties capable of manipulating electromagnetic waves: by blocking, absorbing, enhancing, or bending waves, to achieve benefits that go beyond what is possible with conventional materials.

<span class="mw-page-title-main">Acousto-optics</span> The study of sound and light interaction

Acousto-optics is a branch of physics that studies the interactions between sound waves and light waves, especially the diffraction of laser light by ultrasound through an ultrasonic grating.

Free spectral range (FSR) is the spacing in optical frequency or wavelength between two successive reflected or transmitted optical intensity maxima or minima of an interferometer or diffractive optical element.

When an electromagnetic wave travels through a medium in which it gets attenuated, it undergoes exponential decay as described by the Beer–Lambert law. However, there are many possible ways to characterize the wave and how quickly it is attenuated. This article describes the mathematical relationships among:

References

  1. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 Hecht, Eugene (2002). Optics. Addison-Wesley. ISBN   978-0-321-18878-6.
  2. 1 2 Attwood, David (1999). Soft X-rays and extreme ultraviolet radiation: principles and applications. Cambridge University Press. p. 60. ISBN   978-0-521-02997-1.
  3. Kinsler, Lawrence E. (2000). Fundamentals of Acoustics . John Wiley. p.  136. ISBN   978-0-471-84789-2.
  4. Young, Thomas (1807). A course of lectures on natural philosophy and the mechanical arts. J. Johnson. p.  413.
  5. Newton, Isaac (1730). Opticks: Or, A Treatise of the Reflections, Refractions, Inflections and Colours of Light. William Innys at the West-End of St. Paul's. p.  247.
  6. Hauksbee, Francis (1710). "A Description of the Apparatus for Making Experiments on the Refractions of Fluids". Philosophical Transactions of the Royal Society of London. 27 (325–336): 207. doi:10.1098/rstl.1710.0015. S2CID   186208526.
  7. Hutton, Charles (1795). Philosophical and mathematical dictionary. p. 299. Archived from the original on 2017-02-22.
  8. von Fraunhofer, Joseph (1817). "Bestimmung des Brechungs und Farbenzerstreuungs Vermogens verschiedener Glasarten" [Determination of the Refraction and Color Scattering Power of Different Types of Glass]. Denkschriften der Königlichen Akademie der Wissenschaften zu München[Journal of the Royal Academy of Sciences in Munich] (in German). 5: 208. Archived from the original on 2017-02-22.Exponent des Brechungsverhältnisses is index of refraction
  9. Brewster, David (1815). "On the structure of doubly refracting crystals". Philosophical Magazine. 45 (202): 126. doi:10.1080/14786441508638398. Archived from the original on 2017-02-22.
  10. Herschel, John F. W. (1828). On the Theory of Light. p. 368. Archived from the original on 2015-11-24.
  11. Malitson (1965). "Refractive Index Database". refractiveindex.info. Retrieved June 20, 2018.
  12. Faick, C. A.; Finn, A. N. (July 1931). "The Index of Refraction of Some Soda-Lime-Silica Glasses as a Function of the Composition" (PDF). National Institute of Standards and Technology. Archived (PDF) from the original on December 30, 2016. Retrieved 11 December 2016.
  13. Sultanova, N.; Kasarova, S.; Nikolov, I. (October 2009). "Dispersion Properties of Optical Polymers". Acta Physica Polonica A. 116 (4): 585–587. Bibcode:2009AcPPA.116..585S. doi: 10.12693/APhysPolA.116.585 .
  14. Tapping, J.; Reilly, M. L. (1 May 1986). "Index of refraction of sapphire between 24 and 1060°C for wavelengths of 633 and 799 nm". Journal of the Optical Society of America A. 3 (5): 610. Bibcode:1986JOSAA...3..610T. doi:10.1364/JOSAA.3.000610.
  15. "Forensic Science Communications, Glass Refractive Index Determination". FBI Laboratory Services. Archived from the original on 2014-09-10. Retrieved 2014-09-08.
  16. Tabata, M.; et al. (2005). Development of Silica Aerogel with Any Density (PDF). IEEE Nuclear Science Symposium Conference Record. Vol. 2. pp. 816–818. doi:10.1109/NSSMIC.2005.1596380. ISBN   978-0-7803-9221-2. S2CID   18187536. Archived from the original (PDF) on 2013-05-18.
  17. Sadayori, Naoki; Hotta, Yuji (2004). "Polycarbodiimide having high index of refraction and production method thereof". US Patent Office. US patent 2004/0158021 A1 via Google Patents.
  18. Tosi, Jeffrey L., article on Common Infrared Optical Materials in the Photonics Handbook, accessed on 2014-09-10
  19. Yue, Zengji; Cai, Boyuan; Wang, Lan; Wang, Xiaolin; Gu, Min (2016-03-01). "Intrinsically core-shell plasmonic dielectric nanostructures with ultrahigh refractive index". Science Advances. 2 (3): e1501536. Bibcode:2016SciA....2E1536Y. doi:10.1126/sciadv.1501536. ISSN   2375-2548. PMC   4820380 . PMID   27051869.
  20. 1 2 Als-Nielsen, J.; McMorrow, D. (2011). Elements of Modern X-ray Physics . Wiley-VCH. p.  25. ISBN   978-0-470-97395-0.
  21. 1 2 Gullikson, Eric. "X-Ray interactions with matter". Optical constants. The Center for X-Ray Optics. Lawrence Berkeley Laboratory. Archived from the original on 2011-08-27. Retrieved 2011-08-30.
  22. Lied, Finn (1967). High Frequency Radio Communications with Emphasis on Polar Problems. The Advisory Group for Aerospace Research and Development. pp. 1–7.
  23. Veselago, V. G. (1968). "The electrodynamics of substances with simultaneously negative values of ε and μ". Soviet Physics Uspekhi . 10 (4): 509–514. Bibcode:1968SvPhU..10..509V. doi:10.1070/PU2003v046n07ABEH001614. S2CID   250862458.
  24. Pendry, J. B.; Schurig, D.; Smith, D. R. (Dec 8, 2009). "Electromagnetic compression apparatus, methods and systems". US Patent Office. US Patent 7629941 via Google Patents.
  25. Shalaev, V. M. (2007). "Optical negative-index metamaterials". Nature Photonics . 1 (1): 41–48. Bibcode:2007NaPho...1...41S. doi:10.1038/nphoton.2006.49. S2CID   170678.
  26. 1 2 Feynman, Richard P. (2011). Mainly Mechanics, Radiation, and Heat. Feynman Lectures on Physics. Vol. 1 (The New Millenium ed.). Basic Books. ISBN   978-0-465-02493-3.
  27. 1 2 Paschotta, Rüdiger. "Chromatic Dispersion". RP Photonics Encyclopedia . Archived from the original on 2015-06-29. Retrieved 2023-08-13.
  28. 1 2 Nave, Carl R. (2000). "Dispersion". HyperPhysics . Department of Physics and Astronomy, Georgia State University. Archived from the original on 2014-09-24. Retrieved 2023-08-13.
  29. Paschotta, Rüdiger. "Sellmeier formula". RP Photonics Encyclopedia . Archived from the original on 2015-03-19. Retrieved 2014-09-08.
  30. Schott Company. "Interactive Abbe Diagram". Schott.com. Retrieved 2023-08-13.
  31. Ohara Corporation. "Optical Properties". Oharacorp.com. Retrieved 2022-08-15.
  32. Hoya Group. "Optical Properties". Hoya Group Optics Division. Retrieved 2023-08-13.
  33. Lentes, Frank-Thomas; Clement, Marc K. Th.; Neuroth, Norbert; Hoffmann, Hans-Jürgen; Hayden, Yuiko T.; Hayden, Joseph S.; Kolberg, Uwe; Wolff, Silke (1998). "Optical Properties". In Bach, Hans; Neuroth, Norbert (eds.). The Properties of Optical Glass. Schott Series on Glass and Glass Ceramics. p. 30. doi:10.1007/978-3-642-57769-7. ISBN   978-3-642-63349-2.
  34. Krey, Stefan; Off, Dennis; Ruprecht, Aiko (2014-03-08). "Measuring the Refractive Index with Precision Goniometers: A Comparative Study". In Soskind, Yakov G.; Olson, Craig (eds.). Proc. SPIE 8992, Photonic Instrumentation Engineering. SPIE OPTO, 2014. Vol. 8992. San Francisco, California: SPIE. pp. 56–65. Bibcode:2014SPIE.8992E..0DK. doi:10.1117/12.2041760. S2CID   120544352.
  35. Rupp, Fabian; Jedamzik, Ralf; Bartelmess, Lothar; Petzold, Uwe (2021-09-12). "The modern way of refractive index measurement of optical glass at SCHOTT". In Völkel, Reinhard; Geyl, Roland; Otaduy, Deitze (eds.). Optical Fabrication, Testing, and Metrology VII. Vol. 11873. SPIE. pp. 15–22. Bibcode:2021SPIE11873E..08R. doi:10.1117/12.2597023. ISBN   9781510645905. S2CID   240561530.{{cite book}}: |journal= ignored (help)
  36. "Abbe Refractometer| ATAGO CO., LTD". www.atago.net. Retrieved 2022-08-15.
  37. "Abbe Multi-Wavelength Refractometer". Nova-Tech International. Retrieved 2022-08-15.
  38. Bach, Hans; Neuroth, Norbert, eds. (1998). The Properties of Optical Glass. Schott Series on Glass and Glass Ceramics. p. 267. doi:10.1007/978-3-642-57769-7. ISBN   978-3-642-63349-2.
  39. Dresselhaus, M. S. (1999). "Solid State Physics Part II Optical Properties of Solids" (PDF). Course 6.732 Solid State Physics. MIT. Archived (PDF) from the original on 2015-07-24. Retrieved 2015-01-05.
  40. R. Paschotta, article on optical thickness Archived 2015-03-22 at the Wayback Machine in the Encyclopedia of Laser Physics and Technology Archived 2015-08-13 at the Wayback Machine , accessed on 2014-09-08
  41. R. Paschotta, article on refraction Archived 2015-06-28 at the Wayback Machine in the Encyclopedia of Laser Physics and Technology Archived 2015-08-13 at the Wayback Machine , accessed on 2014-09-08
  42. 1 2 3 4 Born, Max; Wolf, Emil (1999). Principles of Optics (7th expanded ed.). CUP Archive. p.  22. ISBN   978-0-521-78449-8.
  43. Paschotta, R. "Total Internal Reflection". RP Photonics Encyclopedia. Archived from the original on 2015-06-28. Retrieved 2015-08-16.
  44. Swenson, Jim (November 10, 2009). "Refractive Index of Minerals". Newton BBS / Argonne National Laboratory. US DOE. Archived from the original on May 28, 2010. Retrieved 2010-07-28. Incorporates Public Domain material from the U.S. Department of Energy
  45. Nave, Carl R. "Lens-makers' formula". HyperPhysics. Department of Physics and Astronomy. Georgia State University. Archived from the original on 2014-09-26. Retrieved 2014-09-08.
  46. 1 2 3 Carlsson, Kjell (2007). Light microscopy (PDF) (Report). Archived (PDF) from the original on 2015-04-02. Retrieved 2015-01-02.
  47. Bleaney, B.; Bleaney, B.I. (1976). Electricity and Magnetism (Third ed.). Oxford University Press. ISBN   978-0-19-851141-0.
  48. Andrews, David L. (24 February 2015). Photonics, Volume 2: Nanophotonic Structures and Materials. John Wiley & Sons. p. 54. ISBN   978-1-118-22551-6.
  49. Wooten, Frederick (1972). Optical Properties of Solids. New York City: Academic Press. p. 49. ISBN   978-0-12-763450-0. (online pdf) Archived 2011-10-03 at the Wayback Machine
  50. "Calculation of the Refractive Index of Glasses". Statistical Calculation and Development of Glass Properties. Archived from the original on 2007-10-15.
  51. Stone, Jack A.; Zimmerman, Jay H. (2011-12-28). "Index of refraction of air". Engineering metrology toolbox. National Institute of Standards and Technology (NIST). Archived from the original on 2014-01-11. Retrieved 2014-01-11.
  52. Bor, Z.; Osvay, K.; Rácz, B.; Szabó, G. (1990). "Group refractive index measurement by Michelson interferometer". Optics Communications. 78 (2): 109–112. Bibcode:1990OptCo..78..109B. doi:10.1016/0030-4018(90)90104-2.
  53. Milonni, Peter W.; Boyd, Robert W. (2010-12-31). "Momentum of Light in a Dielectric Medium". Advances in Optics and Photonics. 2 (4): 519. Bibcode:2010AdOP....2..519M. doi:10.1364/AOP.2.000519. ISSN   1943-8206.
  54. Young, A.T. (2011). "Refractivity of Air". Archived from the original on 10 January 2015. Retrieved 31 July 2014.
  55. Barrell, H.; Sears, J.E. (1939). "The refraction and dispersion of air for the visible spectrum". Philosophical Transactions of the Royal Society of London . A, Mathematical and Physical Sciences. 238 (786): 1–64. Bibcode:1939RSPTA.238....1B. doi: 10.1098/rsta.1939.0004 . JSTOR   91351.
  56. Aparicio, Josep M.; Laroche, Stéphane (2 June 2011). "An evaluation of the expression of the atmospheric refractivity for GPS signals". Journal of Geophysical Research . 116 (D11): D11104. Bibcode:2011JGRD..11611104A. doi: 10.1029/2010JD015214 .
  57. Ciddor, P.E. (1996). "Refractive index of air: New equations for the visible and near infrared". Applied Optics. 35 (9): 1566–1573. Bibcode:1996ApOpt..35.1566C. doi:10.1364/ao.35.001566. PMID   21085275.
  58. R. Paschotta, article on birefringence Archived 2015-07-03 at the Wayback Machine in the Encyclopedia of Laser Physics and Technology Archived 2015-08-13 at the Wayback Machine , accessed on 2014-09-09
  59. "The Evolution of the Abbe Refractometer". Humboldt State University, Richard A. Paselk. 1998. Archived from the original on 2011-06-12. Retrieved 2011-09-03.
  60. "Refractometers and refractometry". Refractometer.pl. 2011. Archived from the original on 2011-10-20. Retrieved 2011-09-03.
  61. "Refractometer". The Gemology Project. Archived from the original on 2011-09-10. Retrieved 2011-09-03.
  62. Fitzgerald, Richard (July 2000). "Phase‐Sensitive X‐Ray Imaging". Physics Today. 53 (7): 23. Bibcode:2000PhT....53g..23F. doi: 10.1063/1.1292471 . S2CID   121322301.