Diffuse reflectance spectroscopy

Last updated

Diffuse reflectance spectroscopy, or diffuse reflection spectroscopy, is a subset of absorption spectroscopy. It is sometimes called remission spectroscopy. Remission is the reflection or back-scattering of light by a material, while transmission is the passage of light through a material. The word remission implies a direction of scatter, independent of the scattering process. Remission includes both specular and diffusely back-scattered light. The word reflection often implies a particular physical process, such as specular reflection.

Contents

The use of the term remission spectroscopy is relatively recent, and found first use in applications related to medicine and biochemistry. While the term is becoming more common in certain areas of absorption spectroscopy, the term diffuse reflectance is firmly entrenched, as in diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) and diffuse-reflectance ultraviolet–visible spectroscopy.

The mathematical treatments of absorption spectroscopy for scattering materials were originally largely borrowed from other fields. The most successful treatments use the concept of dividing a sample into layers, called plane parallel layers. They are generally those consistent with a two-flux or two-stream approximation. Some of the treatments require all the scattered light, both remitted and transmitted light, to be measured. Others apply only to remitted light, with the assumption that the sample is "infinitely thick" and transmits no light. These are special cases of the more general treatments.

There are several general treatments, all of which are compatible with each other, related to the mathematics of plane parallel layers. They are the Stokes formulas, [1] equations of Benford, [2] Hecht finite difference formula, [3] and the Dahm equation. [4] [5] For the special case of infinitesimal layers, the Kubelka–Munk [6] and Schuster–Kortüm [7] [8] treatments also give compatible results. Treatments which involve different assumptions and which yield incompatible results are the Giovanelli [9] exact solutions, and the particle theories of Melamed [10] and Simmons. [11]

George Gabriel Stokes

George Gabriel Stokes (not to neglect the later work of Gustav Kirchhoff) is often given credit for having first enunciated the fundamental principles of spectroscopy. In 1862, Stokes published formulas for determining the quantities of light remitted and transmitted from "a pile of plates". He described his work as addressing a "mathematical problem of some interest". He solved the problem using summations of geometric series, but the results are expressed as continuous functions. This means that the results can be applied to fractional numbers of plates, though they have the intended meaning only for an integral number. The results below are presented in a form compatible with discontinuous functions.

Stokes used the term "reflexion", not "remission", specifically referring to what is often called regular or specular reflection. In regular reflection, the Fresnel equations describe the physics, which includes both reflection and refraction, at the optical boundary of a plate. "A pile of plates" is still a term of art used to describe a polarizer in which a polarized beam is obtained by tilting a pile of plates at an angle to an unpolarized incident beam. The area of polarization was specifically what interested Stokes in this mathematical problem.

Stokes formulas for remission from and transmission through a "pile of plates"

For a sample that consists of n layers, each having its absorption, remission, and transmission (ART) fractions symbolized by {a, r, t } , with a + r + t = 1, one may symbolize the ART fractions for the sample as {Αn, Rn, Tn} and calculate their values by

where

and

Franz Arthur Friedrich Schuster

In 1905, in an article entitled "Radiation through a foggy atmosphere", Arthur Schuster published a solution to the equation of radiative transfer, which describes the propagation of radiation through a medium, affected by absorption, emission, and scattering processes. [12] His mathematics used a two flux approximation; i.e., all light is assumed to travel with a component either in the same direction as the incident beam, or in the opposite direction. He used the word scattering rather than reflection, and considered scatter to be in all directions. He used the symbols k and s for absorption and isotropic scattering coefficients, and repeatedly refers to radiation entering a "layer", which ranges in size from infinitesimal to infinitely thick. In his treatment, the radiation enters the layers at all possible angles, referred to as "diffuse illumination".

Kubelka and Munk

In 1931, Paul Kubelka (with Franz Munk) published "An article on the optics of paint", the contents of which has come to be known as the Kubelka-Munk theory. They used absorption and remission (or back-scatter) constants, noting (as translated by Stephen H. Westin) that "an infinitesimal layer of the coating absorbs and scatters a certain constant portion of all the light passing through it". While symbols and terminology are changed here, it seems clear from their language that the terms in their differential equations stand for absorption and backscatter (remission) fractions. They also noted that the reflectance from an infinite number of these infinitesimal layers is "solely a function of the ratio of the absorption and back-scatter (remission) constants a0/r0, but not in any way on the absolute numerical values of these constants". This turns out to be incorrect for layers of finite thickness, and the equation was modified for spectroscopic purposes (below), but Kubelka-Munk theory has found extensive use in coatings. [13] [14]

However, in revised presentations of their mathematical treatment, including that of Kubelka, Kortüm and Hecht (below), the following symbolism became popular, using coefficients rather than fractions:

The Kubelka–Munk equation

The Kubelka–Munk equation describes the remission from a sample composed of an infinite number of infinitesimal layers, each having a0 as an absorption fraction, and r0 as a remission fraction.

Deane B. Judd

Deane Judd was very interested the effect of light polarization and degree of diffusion on the appearance of objects. He made important contributions to the fields of colorimetry, color discrimination, color order, and color vision. Judd defined the scattering power for a sample as Sd, where d is the particle diameter. This is consistent with the belief that the scattering from a single particle is conceptually more important than the derived coefficients.

The above Kubelka–Munk equation can be resolved for the ratio a0/r0 in terms of R. This led to a very early (perhaps the first) use of the term "remission" in place of "reflectance" when Judd defined a "remission function" as , where k and s are absorption and scattering coefficients, which replace a0 and r0 in the Kubelka–Munk equation above. Judd tabulated the remission function as a function of percent reflectance from an infinitely thick sample. [15] This function, when used as a measure of absorption, was sometimes referred to as "pseudo-absorbance", a term which has been used later with other definitions [16] as well.

General Electric

In the 1920s and 30s, Albert H. Taylor, Arthur C. Hardy , and others of the General Electric company developed a series of instruments that were capable of easily recording spectral data "in reflection". Their display preference for the data was "% Reflectance". In 1946, Frank Benford [2] published a series of parametric equations that gave results equivalent to the Stokes formulas. The formulas used fractions to express reflectance and transmittance.

Equations of Benford

If A1, R1, and T1 are known for the representative layer of a sample, and An, Rn and Tn are known for a layer composed of n representative layers, the ART fractions for a layer with thickness of n + 1 are

If Ad, Rd and Td are known for a layer with thickness d, the ART fractions for a layer with thickness of d/2 are

and the fractions for a layer with thickness of 2d are

If Ax, Rx and Tx are known for layer x and AyRy and Ty are known for layer y, the ART fractions for a sample composed of layer x and layer y are

The symbol refers to the reflectance of layer when the direction of illumination is antiparallel to that of the incident beam. The difference in direction is important when dealing with inhomogeneous layers. This consideration was added by Paul Kubelka [17] in 1954.

Giovanelli and Chandrasekhar

In 1955, Ron Giovanelli published explicit expressions for several cases of interest which are touted as exact solutions to the radiative transfer equation for a semi-infinite ideal diffuser. [9] His solutions have become the standard against which results from approximate theoretical treatments are measured. Many of the solutions appear deceptively simple due to the work of Subrahmanyan (Chandra) Chandrasekhar. For example, the total reflectance for light incident in the direction μ0 is

Here ω0 is known as the albedo of single scatter σ/(α+σ), representing the fraction of the radiation lost by scattering in a medium where both absorption (α) and scattering (σ) take place. The function H0) is called the H-integral, the values of which were tabulated by Chandrasekhar. [18]

Gustav Kortüm

Kortüm was a physical chemist who had a broad range of interests, and published prolifically. His research covered many aspects of light scattering. He began to pull together what was known in various fields into an understanding of how “reflectance spectroscopy” worked. In 1969, the English translation of his book entitled Reflectance Spectroscopy (long in preparation and translation) was published. This book came to dominate thinking of the day for 20 years in the emerging fields of both DRIFTS and NIR Spectroscopy.

Kortüm's position was that since regular (or specular) reflection is governed by different laws than diffuse reflection, they should therefore be accorded different mathematical treatments. He developed an approach based on Schuster's work by ignoring the emissivity of the clouds in the "foggy atmosphere". If we take α as the fraction of incident light absorbed and σ as the fraction scattered isotropically by a single particle (referred to by Kortüm as the "true coefficients of single scatter"), and define the absorption and isotropic scattering for a layer as and then:

This is the same "remission function" as used by Judd, but Kortüm's translator referred to it as "the so-called reflectance function". If we substitute back for the particle properties, we obtain and then we obtain the Schuster equation for isotropic scattering:

Additionally, Kortüm derived "the Kubelka-Munk exponential solution" by defining k and s as the absorption and scattering coefficient per centimeter of the material and substituting: K ≡ 2k and S ≡ 2s, while pointing out in a footnote that S is a back-scattering coefficient. He wound up with what he called the "Kubelka–Munk function", commonly called the Kubelka–Munk equation:

Kortüm concluded that "the two constant theory of Kubelka and Munk leads to conclusions accessible to experimental test. In practice these are found to be at least qualitatively confirmed, and suitable conditions fulfilling the assumptions made, quantitatively as well."

Kortüm tended to eschew the "particle theories", though he did record that one author, N.T. Melamed of Westinghouse Research Labs, "abandoned the idea of plane parallel layers and substituted them with a statistical summation over individual particles." [19]

Hecht and Simmons

In 1966, Harry G. Hecht (with Wesley W. Wendlandt) published a book entitled "Reflectance Spectroscopy", because "unlike transmittance spectroscopy, there were no reference books written on the subject" of "diffuse reflectance spectroscopy", and "the fundamentals were only to be found in the old literature, some of which was not readily accessible". [20] Hecht describes himself as a novice in the field at the time, and said that if he had known that Gustav Kortüm "a great pillar in the field" was in the process of writing a book on the subject, he "would not have undertaken the task". [21] Hecht was asked to write a review of Kortüm's book [8] and their correspondence concerning it led to Hecht spending a year in Kortüm's laboratories. Kortüm is the author most often cited in the book.

One of the features of the remission function emphasized by Hecht was the fact that

should yield the absorption spectrum displaced by -log s. While the scattering coefficient might change with particle size, the absorption coefficient, which should be proportional to concentration of an absorber, would be obtainable by a background correction for a spectrum. However, experimental data showed the relationship did not hold in strongly absorbing materials. Many papers were published with various explanations for this failure of the Kubelka-Munk equation. Proposed culprits included: incomplete diffusion, anisotropic scatter ("the invalid assumption that radiation is returned equally in all directions from a given particle"), and presence of regular reflection. The situation resulted in a myriad of models and theories being proposed to correct these supposed deficiencies. The various alternative theories were evaluated and compared. [3] [22]

In his book, Hecht reported the mathematics of Stokes and Melamed formulas (which he called “statistical methods”). He believed the approach of Melamed, [19] which “involve a summation over individual particles” was more satisfactory than summations over “plane parallel layers”. Unfortunately, Melamed's method failed as the refractive index of the particles approached unity, but he did call attention to the importance of using individual particle properties, as opposed to coefficients that represent averaged properties for a sample. E.L. Simmons used a simplified modification of the particle model to relate diffuse reflectance to fundamental optical constants without the use of the cumbersome equations. In 1975, Simmons evaluated various theories of diffuse reflectance spectroscopy and concluded that a modified particle model theory is probably the most nearly correct.

In 1976, Hecht wrote a lengthy paper comprehensively describing the myriad of mathematical treatments that had been proposed to deal with diffuse reflectance. In this paper, Hecht states that he assumed (as did Simmons) that in the plane-parallel treatment, the layers could not be made infinitesimally small, but should be restricted to layers of finite thickness interpreted as the mean particle diameter of the sample. This is also supported by the observation that the ratio of the Kubelka–Munk absorption and scattering coefficients is 38 that of corresponding ratio of the Mie coefficients for a sphere. That factor can be rationalized by simple geometric considerations, [5] recognizing that to a first approximation, the absorption is proportional to volume and the scatter is proportional to cross sectional surface area. This is entirely consistent with the Mie coefficients measuring absorption and scatter at a point, and the Kubelka–Munk coefficients measuring scatter by a sphere.

To correct this deficiency of the Kubelka–Munk approach, for the case of an infinitely thick sample, Hecht blended the particle and layer methods by replacing the differential equations in the Kubelka–Munk treatment by finite difference equations, and obtained the Hecht finite difference formula:

Hecht apparently did not know that this result could be generalized, but he realized that the above formula "represents an improvement … and shows the need to consider the particulate nature of scattering media in developing a more precise theory". [3]

Karl Norris (USDA), Gerald Birth

Karl Norris pioneered the field of near-infrared spectroscopy. [23] He began by using log(1/R) as a metric of absorption. While often the samples examined were “infinitely thick”, partially transparent samples were analyzed (especially later) in cells that had a rear reflecting surface (reflector) in a mode called "transflectance". Therefore, the remission from the sample contained light that was back-scattered from the sample, as well as light that was transmitted through the sample, then reflected back to be transmitted through the sample again, thereby doubling the path length. Having no sound theoretical basis for data treatment, Norris used the same electronic processing that was used for absorption data collected in transmission. [24] He pioneered the use of multiple linear regression for analysis of data.

Gerry Birth was the founder of the International Diffuse Reflectance Conference (IDRC). He also worked at the USDA. He was known to have a deep desire to have a better understanding of the process of light scattering. He teamed up with Harry Hecht (who was active in the early meetings of IDRC) to write the Physics theory chapter, with many photographic illustrations, in an influential Handbook edited by Phil Williams and Karl Norris: [25] Nearinfrared Technology in the Agriculture and Food Industries.

Donald J. Dahm, Kevin D. Dahm

In 1994, Donald and Kevin Dahm began using numerical techniques to calculate remission and transmission from samples of varying numbers of plane parallel layers from absorption and remission fractions for a single layer. Their plan was to "start with a simple model, treat the problem numerically rather than analytically, then look for analytical functions that describe the numerical results. Assuming success with that, the model would be made more complex, allowing more complex analytical expressions to be derived, eventually, leading to an understanding of diffuse reflection at a level that appropriately approximated particulate samples." [21] They were able to show the fraction of incident light remitted, R, and transmitted, T, by a sample composed of layers, each absorbing a fraction and remitting a fraction of the light incident upon it, could be quantified by an Absorption/Remission function (symbolized A(R,T) and called the ART function), which is constant for a sample composed of any number of identical layers.

Dahm equation

Also from this process came results for several special cases of two stream solutions for plane parallel layers.

For the case of zero absorption, .

For the case of infinitesimal layers, , and the ART function gives results approaching equivalence to the remission function.

As the void fraction v0 of a layer becomes large, .

The ART is related to the Kortüm–Schuster equation for isotopic scatter by .

The Dahms argued that the conventional absorption and scattering coefficients, as well as the differential equations which employ them, implicitly assume that a sample is homogenous at the molecular level. While this is a good approximation for absorption, as the domain of absorption is molecular, the domain of scattering is the particle as a whole. Any approach using continuous mathematics will therefore tend to fail as particles become large. [26]

Successful application of theory to a real world sample using the mathematics of plane parallel layers requires assigning properties to the layers that are representative of the sample as a whole (which does not require extensively reworking the mathematics). Such a layer was termed a representative layer, and the theory was termed the representative layer theory. [4]

Furthermore, they argued that it was irrelevant whether the light moving from one layer to another was reflected specularly or diffusely. The reflection and back scatter is lumped together as remission. All light leaving the sample on the same side as the incident beam is termed remission, whether it arises from reflection or back scatter. All light leaving the sample on the opposite side from the incident beam is termed transmission. (In a three-flux or higher treatment, such as Giovanelli's, the forward scatter is not indistinguishable from the directly transmitted light. Additionally, Giovanelli's treatment makes the implied assumption of infinitesimal particles.)

They developed a scheme, subject to the limitations of a two-flux model, to calculate the "scatter corrected absorbance" for a sample. [27] The decadic absorbance of a scattering sample is defined as −log10(R+T) or −log10(1−A). For a non scattering sample, R = 0, and the expression becomes −log10T or log(1/T), which is more familiar. In a non-scattering sample, the absorbance has the property that the numerical value is proportional to sample thickness. Consequently, a scatter-corrected absorbance might reasonably be defined as one that has that property.

If one has measured the remission and transmission fractions for a sample, Rs and Ts, then the scatter-corrected absorbance should have half the value for half the sample thickness. By calculating the values for R and T for successively thinner samples (s, 1/2s, 1/4s, …) using the Benford's equations for half thickness, a place will be reached where, for successive values of n (0,1,2,3,...), the expression 2n (−log(R+T)) becomes constant to within a some specified limit, typically 0.01 absorbance units. This value is the scatter-corrected absorbance.

Definitions

Remission

In spectroscopy, remission refers to the reflection or back-scattering of light by a material. While seeming similar to the word "re-emission", it is the light which is scattered back from a material, as opposed to that which is "transmitted" through the material. The word "re-emission" connotes no such directional character. Based on the origin of the word "emit", which means "to send out or away", "re-emit" means "to send out again", "transmit" means "to send across or through", and "remit" means "to send back".

Plane-parallel layers

In spectroscopy, the term "plane parallel layers" may be employed as a mathematical construct in discussing theory. The layers are considered to be semi-infinite. (In mathematics, semi-infinite objects are objects which are infinite or unbounded in some, but not all, possible ways.) Generally, a semi-infinite layer is envisioned as a being bounded by two flat parallel planes, each extending indefinitely, and normal (perpendicular) to the direction of a collimated (or directed) incident beam. The planes are not necessarily physical surfaces which refract and reflect light, but may just describe a mathematical plane, suspended in space. When the plane parallel layers have surfaces, they have been variously called plates, sheets, or slabs.

Representative layer

The term "representative layer" refers to a hypothetical plane parallel layer that has properties relevant to absorption spectroscopy that are representative of a sample as a whole. For particulate samples, a layer is representative if each type of particle in the sample makes up the same fraction of volume and surface area in the layer as in the sample. The void fraction in the layer is also the same as in the sample. Implicit in the representative layer theory is that absorption occurs at the molecular level, but that scatter is from a whole particle.

List of principal symbols used

Note: Where a given letter is used in both capital and lower case form (r, R and t ,T) the capital letter refers to the macroscopic observable and the lower case letter to the corresponding variable for an individual particle or layer of the material. Greek symbols are used for properties of a single particle.

a – absorption fraction of a single layer
r – remission fraction of a single layer
t – transmission fraction of a single layer
An, Rn, Tn – the absorption, remission, and transmission fractions for a sample composed of n layers
α – absorption fraction of a particle
β – back-scattering from a particle
σ – isotropic scattering from a particle
k – absorption coefficient defined as the fraction of incident light absorbed by a very thin layer divided by the thickness of that layer
s – scattering coefficient defined as the fraction of incident light scattered by a very thin layer divided by the thickness of that layer

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 when some kind of radiant excitation intersects a localized phenomenon. 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">Rayleigh scattering</span> Scattering of electromagnetic radiation by particles smaller than the radiations wavelength

Rayleigh scattering, named after the 19th-century British physicist Lord Rayleigh, is the predominantly elastic scattering of light or other electromagnetic radiation by particles much smaller than the wavelength of the radiation. For light frequencies well below the resonance frequency of the scattering particle, the amount of scattering is inversely proportional to the fourth power of the wavelength.

<span class="mw-page-title-main">Reflectance</span> Capacity of an object to reflect light

The reflectance of the surface of a material is its effectiveness in reflecting radiant energy. It is the fraction of incident electromagnetic power that is reflected at the boundary. Reflectance is a component of the response of the electronic structure of the material to the electromagnetic field of light, and is in general a function of the frequency, or wavelength, of the light, its polarization, and the angle of incidence. The dependence of reflectance on the wavelength is called a reflectance spectrum or spectral reflectance curve.

<span class="mw-page-title-main">Scattering</span> Range of physical processes

Scattering is a term used in physics to describe a wide range of physical processes where moving particles or radiation of some form, such as light or sound, are forced to deviate from a straight trajectory by localized non-uniformities in the medium through which they pass. In conventional use, this also includes deviation of reflected radiation from the angle predicted by the law of reflection. Reflections of radiation that undergo scattering are often called diffuse reflections and unscattered reflections are called specular (mirror-like) reflections. Originally, the term was confined to light scattering. As more "ray"-like phenomena were discovered, the idea of scattering was extended to them, so that William Herschel could refer to the scattering of "heat rays" in 1800. John Tyndall, a pioneer in light scattering research, noted the connection between light scattering and acoustic scattering in the 1870s. Near the end of the 19th century, the scattering of cathode rays and X-rays was observed and discussed. With the discovery of subatomic particles and the development of quantum theory in the 20th century, the sense of the term became broader as it was recognized that the same mathematical frameworks used in light scattering could be applied to many other phenomena.

In physics, mean free path is the average distance over which a moving particle travels before substantially changing its direction or energy, typically as a result of one or more successive collisions with other particles.

<span class="mw-page-title-main">Mie scattering</span> Scattering of an electromagnetic wave

The Mie solution to Maxwell's equations describes the scattering of an electromagnetic plane wave by a homogeneous sphere. The solution takes the form of an infinite series of spherical multipole partial waves. It is named after Gustav Mie.

Absorbance is defined as "the logarithm of the ratio of incident to transmitted radiant power through a sample ". Alternatively, for samples which scatter light, absorbance may be defined as "the negative logarithm of one minus absorptance, as measured on a uniform sample". The term is used in many technical areas to quantify the results of an experimental measurement. While the term has its origin in quantifying the absorption of light, it is often entangled with quantification of light which is “lost” to a detector system through other mechanisms. What these uses of the term tend to have in common is that they refer to a logarithm of the ratio of a quantity of light incident on a sample or material to that which is detected after the light has interacted with the sample. 

<span class="mw-page-title-main">Diffuse reflection</span> Reflection with light scattered at random angles

Diffuse reflection is the reflection of light or other waves or particles from a surface such that a ray incident on the surface is scattered at many angles rather than at just one angle as in the case of specular reflection. An ideal diffuse reflecting surface is said to exhibit Lambertian reflection, meaning that there is equal luminance when viewed from all directions lying in the half-space adjacent to the surface.

Cavity ring-down spectroscopy (CRDS) is a highly sensitive optical spectroscopic technique that enables measurement of absolute optical extinction by samples that scatter and absorb light. It has been widely used to study gaseous samples which absorb light at specific wavelengths, and in turn to determine mole fractions down to the parts per trillion level. The technique is also known as cavity ring-down laser absorption spectroscopy (CRLAS).

Fluorescence correlation spectroscopy (FCS) is a statistical analysis, via time correlation, of stationary fluctuations of the fluorescence intensity. Its theoretical underpinning originated from L. Onsager's regression hypothesis. The analysis provides kinetic parameters of the physical processes underlying the fluctuations. One of the interesting applications of this is an analysis of the concentration fluctuations of fluorescent particles (molecules) in solution. In this application, the fluorescence emitted from a very tiny space in solution containing a small number of fluorescent particles (molecules) is observed. The fluorescence intensity is fluctuating due to Brownian motion of the particles. In other words, the number of the particles in the sub-space defined by the optical system is randomly changing around the average number. The analysis gives the average number of fluorescent particles and average diffusion time, when the particle is passing through the space. Eventually, both the concentration and size of the particle (molecule) are determined. Both parameters are important in biochemical research, biophysics, and chemistry.

<span class="mw-page-title-main">Opacity (optics)</span> Property of an object or substance that is impervious to light

Opacity is the measure of impenetrability to electromagnetic or other kinds of radiation, especially visible light. In radiative transfer, it describes the absorption and scattering of radiation in a medium, such as a plasma, dielectric, shielding material, glass, etc. An opaque object is neither transparent nor translucent. When light strikes an interface between two substances, in general some may be reflected, some absorbed, some scattered, and the rest transmitted. Reflection can be diffuse, for example light reflecting off a white wall, or specular, for example light reflecting off a mirror. An opaque substance transmits no light, and therefore reflects, scatters, or absorbs all of it. Other categories of visual appearance, related to the perception of regular or diffuse reflection and transmission of light, have been organized under the concept of cesia in an order system with three variables, including opacity, transparency and translucency among the involved aspects. Both mirrors and carbon black are opaque. Opacity depends on the frequency of the light being considered. For instance, some kinds of glass, while transparent in the visual range, are largely opaque to ultraviolet light. More extreme frequency-dependence is visible in the absorption lines of cold gases. Opacity can be quantified in many ways; for example, see the article mathematical descriptions of opacity.

<span class="mw-page-title-main">Dynamic light scattering</span> Technique for determining size distribution of particles

Dynamic light scattering (DLS) is a technique in physics that can be used to determine the size distribution profile of small particles in suspension or polymers in solution. In the scope of DLS, temporal fluctuations are usually analyzed using the intensity or photon auto-correlation function. In the time domain analysis, the autocorrelation function (ACF) usually decays starting from zero delay time, and faster dynamics due to smaller particles lead to faster decorrelation of scattered intensity trace. It has been shown that the intensity ACF is the Fourier transform of the power spectrum, and therefore the DLS measurements can be equally well performed in the spectral domain. DLS can also be used to probe the behavior of complex fluids such as concentrated polymer solutions.

The linear attenuation coefficient, attenuation coefficient, or narrow-beam attenuation coefficient characterizes how easily a volume of material can be penetrated by a beam of light, sound, particles, or other energy or matter. A coefficient value that is large represents a beam becoming 'attenuated' as it passes through a given medium, while a small value represents that the medium had little effect on loss. The SI unit of attenuation coefficient is the reciprocal metre (m−1). Extinction coefficient is another term for this quantity, often used in meteorology and climatology. Most commonly, the quantity measures the exponential decay of intensity, that is, the value of downward e-folding distance of the original intensity as the energy of the intensity passes through a unit thickness of material, so that an attenuation coefficient of 1 m−1 means that after passing through 1 metre, the radiation will be reduced by a factor of e, and for material with a coefficient of 2 m−1, it will be reduced twice by e, or e2. Other measures may use a different factor than e, such as the decadic attenuation coefficient below. The broad-beam attenuation coefficient counts forward-scattered radiation as transmitted rather than attenuated, and is more applicable to radiation shielding.

<span class="mw-page-title-main">Radiative transfer equation and diffusion theory for photon transport in biological tissue</span>

Photon transport in biological tissue can be equivalently modeled numerically with Monte Carlo simulations or analytically by the radiative transfer equation (RTE). However, the RTE is difficult to solve without introducing approximations. A common approximation summarized here is the diffusion approximation. Overall, solutions to the diffusion equation for photon transport are more computationally efficient, but less accurate than Monte Carlo simulations.

Within physics, the Hybrid Theory for photon transport in tissue uses the advantages and eliminates the deficiencies of both the Monte Carlo method and the diffusion theory for photon transport to model photons traveling through tissue both accurately and efficiently.

Diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) is an infrared spectroscopy sampling technique used on powder samples without prior preparation. The sample is added to a sample cup and the data is collected on the bulk sample. The infrared light on a sample is reflected and transmitted at different amounts depending on the bulk properties of the material. Diffuse reflection of the incident light produced by the sample's rough surface reflection in all directions is collected by use of an ellipsoid or paraboloid mirror. Shape, compactness, refractive index, reflectivity and absorption of the particles are all characteristic of the material being analyzed. If the sample is too absorbent, then it can be diluted with a nonabsorbent material such as potassium bromide, potassium chloride, etc. The particle size should be smaller than the wavelength of the incident light in order to minimize Mie scattering, so this would infer that it should be less than 5 µm for mid-infrared spectroscopy. The spectra are plotted in units of log inverse reflectance versus wavenumber. Alternative plots of Kubelka-Munk units can be used, which relate reflectance to concentration using a scaling factor. A reflectance standard is needed in order to quantify the reflectance of the sample because it cannot be determined directly.

The Kubelka-Munk theory, devised by Paul Kubelka and Franz Munk, is a fundamental approach to modelling the appearance of paint films. As published in 1931, the theory addresses "the question of how the color of a substrate is changed by the application of a coat of paint of specified composition and thickness, and especially the thickness of paint needed to obscure the substrate". The mathematical relationship involves just two paint-dependent constants.

The concept of the representative layer came about though the work of Donald Dahm, with the assistance of Kevin Dahm and Karl Norris, to describe spectroscopic properties of particulate samples, especially as applied to near-infrared spectroscopy. A representative layer has the same void fraction as the sample it represents and each particle type in the sample has the same volume fraction and surface area fraction as does the sample as a whole. The spectroscopic properties of a representative layer can be derived from the spectroscopic properties of particles, which may be determined by a wide variety of ways. While a representative layer could be used in any theory that relies on the mathematics of plane parallel layers, there is a set of definitions and mathematics, some old and some new, which have become part of representative layer theory.

<span class="mw-page-title-main">Hiding power</span> Property of paint

The hiding power is an ability of a paint to hide the surface that the paint was applied to. Numerically, it is defined as an area of surface coated by a volume of paint at which the "complete hiding" of the underlying surface occurs.

References

  1. Stokes, George (1862). "On the intensity of the light reflected from or transmitted through a pile of plates". Proceedings of the Royal Society of London. 11: 545–556. doi: 10.1098/rspl.1860.0119 .
  2. 1 2 Benford, Frank (1946). "Radiation in a Diffusing Medium". Journal of the Optical Society of America. 36 (9): 524–554. doi:10.1364/JOSA.36.000524. PMID   21002043.
  3. 1 2 3 Hecht, Harry (1976). "The Interpretation of Diffuse Reflectance Spectra". Journal of Research of the National Bureau of Standards Section A. 80A (4): 567–583. doi: 10.6028/jres.080A.056 . PMC   5293523 . PMID   32196278.
  4. 1 2 Dahm, Donald (1999). "Representative Layer Theory for Diffuse Reflectance". Applied Spectroscopy. 53 (6): 647–654. Bibcode:1999ApSpe..53..647D. doi:10.1366/0003702991947298. S2CID   96885077.
  5. 1 2 Griffiths, Peter; Dahm, Donald J. (2007). "Continuum and Discontinuum Theories of Diffuse Reflection". In Burns, Donald A. (ed.). Handbook of Near-Infrared Analysis (3rd ed.). Boca Raton: CRC Press. ISBN   9780849373930.
  6. Kubelka, Paul (1931). "Ein Beitrag zur Optik der Farbanstriche" (PDF). Zeits. F. Techn. Physik. 12: 593–601.
  7. Schuster, Aurhur (1905). "Radiation through a foggy atmosphere". Astrophysical Journal. 21 (1): 1–22. Bibcode:1905ApJ....21....1S. doi:10.1086/141186.
  8. 1 2 Kortüm, Gustav (1969). Reflectance spectroscopy Principles, methods, applications. Berlin: Springer. ISBN   9783642880711. OCLC   714802320.
  9. 1 2 Giovanelli, Ronald (1955). "Reflection by semi-infinite diffusers". Optica Acta. 2 (4): 153–162. Bibcode:1955AcOpt...2..153G. doi:10.1080/713821040.
  10. Melamed, N T (1963). "Optical properties of powders. Part I. Optical absorption coefficients and the absolute value of the diffuse reflectance". Journal of Applied Physics. 34: 560. doi:10.1063/1.1729309.
  11. Simmons, E L (1975). "Modification of the particle−model theory of diffuse reflectance properties of powdered samples". Journal of Applied Physics. 46 (1): 344. Bibcode:1975JAP....46..344S. doi:10.1063/1.321341.
  12. Schuster, Arthur (1905-01-01). "Radiation Through a Foggy Atmosphere". The Astrophysical Journal. 21: 1. Bibcode:1905ApJ....21....1S. doi:10.1086/141186. ISSN   0004-637X.
  13. Shen, Jing; Li, Ya; He, Ji-Huan (2016-04-01). "On the Kubelka–Munk absorption coefficient". Dyes and Pigments. 127: 187–188. doi:10.1016/j.dyepig.2015.11.029. ISSN   0143-7208.
  14. Bullett, T.R. (1999), "Appearance qualities of paint — Basic concepts", Paint and Surface Coatings, Elsevier, pp. 621–641, retrieved 2023-06-23
  15. Judd, D B (1963). Color in Business, Science, and Industry (2 ed.). New York: John Wiley & Sons, Inc.
  16. Reeves, James B.; McCarty, Gregory W.; Rutherford, David W.; Wershaw, Robert L. (1 October 2007). "Near Infrared Spectroscopic Examination of Charred Pine Wood, Bark, Cellulose and Lignin: Implications for the Quantitative Determination of Charcoal in Soils". Journal of Near Infrared Spectroscopy. 15 (5): 307–315. Bibcode:2007JNIS...15..307R. doi:10.1255/jnirs.742. ISSN   0967-0335. S2CID   95231598.
  17. Kubelka, Paul (1954-04-01). "New Contributions to the Optics of Intensely Light-Scattering Materials. Part II: Nonhomogeneous Layers*". JOSA. 44 (4): 330–335. doi:10.1364/JOSA.44.000330.
  18. Chandrasekhar, S (1960). Radiative Transfer. New York: Dover Publications, Inc. ISBN   978-0486605906.
  19. 1 2 Melamed, N T (1963). "Optical properties of powders. Part I. Optical absorption coefficients and the absolute value of the diffuse reflectance". Journal of Applied Physics. 34: 560. doi:10.1063/1.1729309.
  20. Wendlandt, Wesley (1969). Reflectance Spectroscopy. New York: Interscience Publishers.
  21. 1 2 Dahm, Donald (2007). Interpreting Diffuse Reflectance and Transmittance: A Theoretical Introduction to Absorption Spectroscopy of Scattering Materials. Chichester, UK: NIR Publications. ISBN   9781901019056.
  22. Simmons, E L (1975). "Diffuse reflectance spectroscopy: a comparison of the theories". Applied Optics. 14 (6): 1380–1386. Bibcode:1975ApOpt..14.1380S. doi:10.1364/AO.14.001380. PMID   20154834.
  23. Williams, Phil (2019-10-01). "Karl H. Norris, the Father of Near-Infrared Spectroscopy". NIR News. 30 (7–8): 25–27. doi: 10.1177/0960336019875883 . ISSN   0960-3360.
  24. Norris, Karl (2005). "Why log(1/R) for Composition Analysis with NIR?". NIR News. 16 (8): 10–13. doi:10.1255/nirn.865. S2CID   100866871.
  25. Birth, Gerald (1983). "The Physics of Near–InfraRed Reflectance", in Nearinfrared Technology in the Agriculture and Food Industries, Ed by Phil Williams and Karl Norris (1 ed.). St Paul, MN: The American Association of Cereal Chemists. ISBN   1891127241.
  26. Dahm, Donald (2003). "Illustration of Failure of Continuum Models of Diffuse Reflectance". Journal of Near Infrared Spectroscopy. 11 (6): 479–485. doi:10.1255/jnirs.398. S2CID   93926306.
  27. Dahm, Kevin (2013). "Separating the Effects of Scatter and Absorption Using the Representative Layer". Journal of Near Infrared Spectroscopy. 21 (5): 351–357. Bibcode:2013JNIS...21..351D. doi:10.1255/jnirs.1062. S2CID   98416407.