Kubelka–Munk theory

Last updated

The Kubelka-Munk theory, [1] [2] devised by Paul Kubelka [3] [4] 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.

Contents

In their article, "fundamental differential equations" are developed using a two-stream approximation for light diffusing through a coating whose absorption and remission (back-scattering) coefficients are known. The total remission from a coating surface is the summation of: 1) the reflectance of the coating surface; 2) the remission from the interior of the coating; and 3) the remission from the surface of the substrate. The intensity considered in the latter two parts is modified by the absorption of the coating material. The concept is based on the simplified picture of two diffuse light fluxes moving through semi-infinite plane-parallel layers, with one flux proceeding "downward", and the other simultaneously "upward".

While Kubelka entered this field through an interest in coatings, his work has influenced workers in other areas as well. In the original article, there is a special case of interest to many fields is "the albedo of an infinitely thick coating". This case yielded the Kubelka–Munk equation, which 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. The authors noted that the remission 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". (The equation is presented in the same mathematical form as in the article, but with symbolism modified. [lower-alpha 1] )

While numerous early authors had developed similar two-constant equations, the mathematics of most of these was found to be consistent with the Kubelka–Munk treatment. [5] Others added additional constants to produce more accurate models, but these generally did not find wide acceptance. Due to its simplicity and its acceptable prediction accuracy in many industrial applications, the Kubelka–Munk model remains very popular. However, in almost every application area, the limitations of the model have required improvements. Sometimes these improvements are touted as extensions of Kubelka–Munk theory, sometimes as embracing more general mathematics of which the Kubelka–Munk equation is a special case, and sometimes as an alternate approach.

Paints

In the original article, [1] there are several special cases important to paints that are addressed, along with a mathematical definition of hiding power (an ability to hide the surface of an object). The hiding power of a coating measures its ability to obscure a background of contrasting color. Hiding power is also known as opacity or covering power.

In the following, R is the fraction of incident light that is remitted (reflected) by a coated substrate under consideration, Rg is the remission fraction from the substrate alone, Rc is the remission fraction from the coating, R is remission fraction of an infinitely thick layer, is the fraction of incident light transmitted by the sample under consideration, and X is the coating thickness.

Ideal white paint
An ideal white paint reflects all incident light, and absorbs none, or and For this case, the remission fraction R for a layer of finite thickness is
Ideal glaze
A coating of an ideal glaze emits no light () and absorbs a fraction For this case, . For an infinitely thick glaze,

In the original article, there is a solution for remission from a coating of finite thickness. Kubelka derived many additional formulas for a variety of other cases, which were published in the post-war years. [6] [7] [8] Whereas the 1931 theory assumed that light flows in one dimension (two fluxes, upward and downward within the layer), in 1948 Kubelka derived the same equations (up to a factor of 2) assuming spherical scatter within the paint layer. Later he generalized the theory to inhomogeneous layers (see below).

Paper and paper coatings

The Kubelka–Munk theory is also used in the paper industry to predict optical properties of paper, avoiding a labor-intensive trial-and-error approach. [9] The theory is relatively simple in terms of the number of constants involved, works very well for many papers, and is well documented for use by the pulp and paper industry. If the optical properties (e.g., reflectance and opacity) of each pulp, filler, and dye used in paper-making are known, then the optical properties of a paper made with any combination of the materials can be predicted. If the contrast ratio and reflectivity of a paper are known, the changes in these properties with a change in basis weight can be predicted. [10]

While the Kubelka–Munk coefficients are assumed to be linear and independent quantities, the relationship fails in regions of strong absorption, such as in the case of dyed paper. Several theories were proposed to explain the non-linear behavior of the coefficients, attributing the non-linearity to the non-isotropic structure of paper at both the micro- and macroscopic levels. [9] However, using an analysis based on the Kramers–Kronig relations, the coefficients were shown to be dependent quantities related to the real and imaginary part of the refractive index. By accounting for this dependency, the anomalous behavior of the Kubelka–Munk coefficients in regions of strong absorption were fully explained. [11]

Semiconductors

The band-gap energy of semiconductors is frequently determined from a Tauc plot, where the quantity is plotted against photon energy E. Then the band-gap energy can be obtained by extending the straight segment of the graph to the E axis. [12] There is a simpler method adapted from the Kubelka–Munk theory, in which the band gap is calculated by plotting, versus E, where is the absorption coefficient. [13]

Color

Early practitioners, especially D. R. Duncan, [14] assumed that in a mixture of pigments, the colors produced in any given medium may be deduced from formulae involving two constants for each pigment. These constants, which vary with the wavelength of the incident light, measure respectively the absorbing power of the pigment for light and its scattering power. The work of Kubelka and Munk was seen as yielding a useful systematic approach to color mixing and matching. By resolving the Kubelka–Munk equation for the ratio of absorption to scatter, one can obtain a "remission function": [15]

We may define and as absorption and back-scattering coefficients, which replace the absorption and remission fractions a0 and r0 in the Kubelka–Munk equation above. Then assuming separate additivity of the absorption and coefficients for each of components of concentration :

For the case of small amount of pigments, the scatter is dominated by the base material and is assumed to be constant. In such a case, the equation is linear in concentration of pigment. [5]

Spectroscopy

One special case has received much attention in diffuse reflectance spectroscopy: [5] that of an opaque (infinitely thick) coating, which can be applied to a sample modeled as an infinite number of infinitesimal layers. The two-stream approximation was embraced by the early practitioners. [16] [17] There were far more mathematics to choose from, but the name Kubelka–Munk became widely regarded as synonymous with any technique that modeled diffuse radiation moving through layers of infinitesimal size. This was aided by the popular assumption that the Kubelka–Munk function (above) was analogous to the absorbance function in transmission spectroscopy.

In the field of infrared spectroscopy, it was common to prepare solid samples by finely grinding the sample with potassium bromide (KBr). This led to a situation analogous to the described in the section just above for pigments, where the analyte had little effect on the scatter, which was dominated by the KBr. In this case, the assumption of the function being linear with concentration was reasonable.

However, in the field of near-infrared spectroscopy, the samples are generally measured in their natural (often particulate) state, and deviations from linearity at higher absorption levels were routinely observed. The remission function (also called the Kubelka–Munk function) was almost abandoned in favor of "log(1/R)". A more general equation, called the Dahm equation, was developed, [18] along with a scheme to separate the effects of scatter from absorption in the log(1/R) data. [19] In the equation, and are the measured remission by and transmission through a sample of layers, each layer having absorption and remission fractions of and R. Note that the so-called ART function

is constant for any sample thickness.

In other areas of spectroscopy, there are shifts away from the strict use of the Kubelka–Munk treatment as well. [20]

Failure of continuous models of diffuse reflectance

Continuous models are widely used to model diffuse reflection from particulate samples. They are embodied in various theories, including diffusion theory, the equation of radiation transfer, as well as Kubelka–Munk. In spite of its widespread use, there has long been an understanding that the Kubelka–Munk (K–M) theory has limitations. The term "failure of the Kubelka–Munk theory" has been applied because it does not "remain valid in strongly absorbing materials". [5] There have been many attempts to explain the limitations and amend the K–M equation. [21] In literature related to diffuse-reflection infrared Fourier-transform (DRIFT) spectra, "particularly specular reflection" is often identified as a culprit. [22] In some corners, there is the working assumption that the problem is that the K–M theory is a two-flux theory, and that introducing additional directions will solve the problem. In particular, two continuous theories, "diffusion theory" and the "equation of radiation transfer" (ERT), have their advocates. Some of the advocates of the ERT have called to our attention the failure of the ERT to predict the desired linear absorption coefficient as particle size gets large, and blamed it on the hidden mass effect. [23] In 2003, [24] Donald and Kevin Dahm illustrated the degree to which the continuous theories all suffer from the fundamental limitation of trying to model a discontinuous sample as a continuum and suggested that as long as the effect of this limitation is unexplored, there is little reason to search for other reasons for "failure".

Spectroscopists have a desire to determine the same absorption coefficient quantity from diffuse reflectance measurements as they would from a transmission measurement on a non-scattering sample of the same material. The Bouguer–Lambert law describes the attenuation of transmitted light as exponential falloff in intensity of a direct beam of light as it passes through a medium. The cause of the attenuation may be absorption or scatter. A coefficient unaffected by scatter is desired by absorption spectroscopists. Mathematically, the Bouguer–Lambert law may be expressed as

where is the linear absorption coefficient, is the back-scatter coefficient, and is their sum, often called the extinction coefficient. (The symbols and may be used to represent the absorption and scattering parts of the extinction coefficients.)

Through work with the Dahm equation, we know that the ART function is constant for all sample thicknesses of the same material. This would include the infinitesimal layer used in the Kubelka–Munk differentiation. Consequently we may equate numerous functions:

Using a simple system (albeit rather complex mathematics), it can be shown that continuous models correctly predict the R and in the ART function, but do not correctly predict the fractions of incident light that are transmitted directly. From this, it can be deduced that the coefficients and and not proportional to the and in the Bouguer–Lambert law.

Treatment of inhomogeneous layers

A coating layer is not the same as the substrate it covers. As Kubelka was interested in coatings, he was of course very interested in the handling of what he called "inhomogeneous layers". A set of equations, one of which was believed to apply to the case, had been published by Frank Benford [25] in 1946 for the case of two light streams through plane parallel layers. However, it did not handle it successfully. Kubelka solved the problem, and we illustrate the solution here. First, a case to which the equation of Benford may be straightforwardly applied.

Kubelka1.jpg

The sketch shows two surfaces bounding a slab of a non-absorbing medium. Notice that the assembly would appear identical regardless of which side was being entered. Apart from the surfaces, the medium has no spectroscopic properties. A beam of light of unit intensity reaches the front surface, and by our assumptions, half is remitted and half is transmitted. The portion that is transmitted proceeds to the other surface undiminished. There it is again split where half (1/4 of the original incident intensity) is transmitted and half is remitted. The amounts that remitted from the first surface can be totaled as can the amount that are transmitted through the second. The total remission is 2/3 ≈ 0.667. The total transmission is 1/3 ≈ 0.333.

Alternatively, we can use the equation of Benford that applies. For two plane parallel layers, x and y, having different properties, the transmission remission , and absorption fractions for the two layers can be calculated from the properties of the individual layers () from the following equations:

Next we will examine the case where the medium is absorbing one. While the total assembly would behave the same in either direction, in order to apply the mathematics, we will need to use an intermediate step where it does not.

Kubelka 2.JPG

Here we will assume that again the surfaces will remit and transmit 1/2 the amount striking it, this time we will assume that half of the intensity will be absorbed a trip across the slab. A beam of light of unit intensity reaches the front surface, and by our assumptions, half is remitted and half is transmitted, but this time half of the transmitted light, or 1/4 is absorbed before another 1/4 reaches the second surface, and 1/4 is remitted back across the slab to face half of it being absorbed. The sketch shows that the calculated values from the equations should be R = 8/15 ≈ 0.533, T = 2/15 ≈ 0.133, A = 1/3 ≈ 0.333.

Now the sketch has three layers, labeled 1, 2, and 3. Layers 1 and 3 remit and transmit 1/2 and absorb nothing. Layer 2 absorbs half and transmits half, but remits nothing. We can build the assembly by first combining layers 1 and 2, and then combining that result as the x value in combing with layer 3 (as y).

So for step 1:

Kubelka [8] has shown by theory and experiment that remittance and absorption of a non-homogeneous specimen depend on the direction of illumination, whereas transmittance does not. Consequently, for non-homogeneous layers, the remission from the first layer that occurs in the denominator is the remission when illuminated from the reverse (not the forward) direction, so we will need to know the value for for the next step, that is when the layer x is layer 2, and layer y is layer 1:

The next step is then to set and , with and , with in the denominator as :

In computer art

The K–M paint-mixing algorithm has been adapted to directly use the RGB color model by Sochorová and Jamriška in 2021. Their "Mixbox" approach works by converting the inputs into a version of CMYK (phthalo blue, quinacridone magenta, Hansa yellow, and titanium white) plus a residue (to account for the gamut difference), performing the K–M mixing in that latent space, and then producing the output in RGB. There are additional concerns for dealing with wider gamuts and improving speed. This RGB adaptation makes it easier for digital painting software to integrate the more realistic K–M method. [26]

Notes

  1. The original symbolism was: , where confusingly is related to absorption, not scatter. The is for the German word Helligkeit, "brightness".

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">Fick's laws of diffusion</span> Mathematical descriptions of molecular diffusion

Fick's laws of diffusion describe diffusion and were first posited by Adolf Fick in 1855 on the basis of largely experimental results. They can be used to solve for the diffusion coefficient, D. Fick's first law can be used to derive his second law which in turn is identical to the diffusion equation.

In mathematics, an algebraic equation or polynomial equation is an equation of the form

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. 

In mathematics and its applications, a Sturm–Liouville problem is a second-order linear ordinary differential equation of the form:

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.

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.

<span class="mw-page-title-main">Monte Carlo method for photon transport</span>

Modeling photon propagation with Monte Carlo methods is a flexible yet rigorous approach to simulate photon transport. In the method, local rules of photon transport are expressed as probability distributions which describe the step size of photon movement between sites of photon-matter interaction and the angles of deflection in a photon's trajectory when a scattering event occurs. This is equivalent to modeling photon transport analytically by the radiative transfer equation (RTE), which describes the motion of photons using a differential equation. However, closed-form solutions of the RTE are often not possible; for some geometries, the diffusion approximation can be used to simplify the RTE, although this, in turn, introduces many inaccuracies, especially near sources and boundaries. In contrast, Monte Carlo simulations can be made arbitrarily accurate by increasing the number of photons traced. For example, see the movie, where a Monte Carlo simulation of a pencil beam incident on a semi-infinite medium models both the initial ballistic photon flow and the later diffuse propagation.

<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.

<span class="mw-page-title-main">Transfer-matrix method (optics)</span>

The transfer-matrix method is a method used in optics and acoustics to analyze the propagation of electromagnetic or acoustic waves through a stratified medium; a stack of thin films. This is, for example, relevant for the design of anti-reflective coatings and dielectric mirrors.

Surface-extended X-ray absorption fine structure (SEXAFS) is the surface-sensitive equivalent of the EXAFS technique. This technique involves the illumination of the sample by high-intensity X-ray beams from a synchrotron and monitoring their photoabsorption by detecting in the intensity of Auger electrons as a function of the incident photon energy. Surface sensitivity is achieved by the interpretation of data depending on the intensity of the Auger electrons instead of looking at the relative absorption of the X-rays as in the parent method, EXAFS.

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.

In engineering, physics, and chemistry, the study of transport phenomena concerns the exchange of mass, energy, charge, momentum and angular momentum between observed and studied systems. While it draws from fields as diverse as continuum mechanics and thermodynamics, it places a heavy emphasis on the commonalities between the topics covered. Mass, momentum, and heat transport all share a very similar mathematical framework, and the parallels between them are exploited in the study of transport phenomena to draw deep mathematical connections that often provide very useful tools in the analysis of one field that are directly derived from the others.

In mathematics, the limiting absorption principle (LAP) is a concept from operator theory and scattering theory that consists of choosing the "correct" resolvent of a linear operator at the essential spectrum based on the behavior of the resolvent near the essential spectrum. The term is often used to indicate that the resolvent, when considered not in the original space (which is usually the space), but in certain weighted spaces (usually , see below), has a limit as the spectral parameter approaches the essential spectrum. This concept developed from the idea of introducing complex parameter into the Helmholtz equation for selecting a particular solution. This idea is credited to Vladimir Ignatowski, who was considering the propagation and absorption of the electromagnetic waves in a wire. It is closely related to the Sommerfeld radiation condition and the limiting amplitude principle (1948). The terminology – both the limiting absorption principle and the limiting amplitude principle – was introduced by Aleksei Sveshnikov.

Time-domain diffuse optics or time-resolved functional near-infrared spectroscopy is a branch of functional near-Infrared spectroscopy which deals with light propagation in diffusive media. There are three main approaches to diffuse optics namely continuous wave (CW), frequency domain (FD) and time-domain (TD). Biological tissue in the range of red to near-infrared wavelengths are transparent to light and can be used to probe deep layers of the tissue thus enabling various in vivo applications and clinical trials.

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. 1 2 Kubelka, Paul; Munk, Franz (1931). "An article on optics of paint layers" (PDF). Z. Tech. Phys. 12: 593–601.
  2. "Kubelka–Munk Theory  an overview". www.sciencedirect.com. Retrieved 22 February 2021.
  3. Brill, Michael; Vik, Michal (2014). "Paul Kubelka".
  4. Westin, Stephen (2011). "Autobiographical Sketch: Paul Kubelka".
  5. 1 2 3 4 Wendlandt, Wesley; Hecht, Harry (1966). Reflectance Spectroscopy. New York: Interscience.
  6. Kubelka, Paul (1948-05-01). "New Contributions to the Optics of Intensely Light-Scattering Materials. Part I". JOSA. 38 (5): 448–457. Bibcode:1948JOSA...38..448K. doi:10.1364/JOSA.38.000448. PMID   18916891.
  7. Kubelka, Paul (1948-12-01). "Errata: New Contributions to the Optics of Intensely Light-Scattering Materials. Part I". JOSA. 38 (12): 1067. doi:10.1364/JOSA.38.001067.
  8. 1 2 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.
  9. 1 2 Hubbe, M. A.; Pawlak, J. J.; Koukoulas, A. A. (2008). "Paper's Appearance: A Review". BioResources . 3 (2): 627–665. doi: 10.15376/biores.3.2.627-665 .
  10. Bajpai, Pratima (2018). "Chapter 11. Optical Properties of Paper". Biermann's Handbook of Pulp and Paper. Vol. 2: Paper and Board Making (3rd ed.). Elsevier Inc. pp. 237–271. doi:10.1016/B978-0-12-814238-7.00011-8. ISBN   9780128142387.
  11. Koukoulas, A. A.; Jordan, B. D. (May 1997). "Effect of strong absorption on the Kubelka-Munk scattering coefficient". Journal of Pulp and Paper Science. 23: J224–J232.
  12. Makuła, Patrycja; Pacia, Michał; Macyk, Wojciech (2018-12-06). "How To Correctly Determine the Band Gap Energy of Modified Semiconductor Photocatalysts Based on UV–Vis Spectra". The Journal of Physical Chemistry Letters. 9 (23): 6814–6817. doi: 10.1021/acs.jpclett.8b02892 . PMID   30990726.
  13. Abdullahi, Sabiu (2016). "Sımple Method For The Determınatıon of Band Gap of a Nanopowdered Sample Usıng Kubelka Munk Theory". Journal of the Nigerian Association of Mathematical Physics. 35: 241–246.
  14. Duncan, D. R. (1940). "The colour of pigment mixtures". Proceedings of the Physical Society. 52 (3): 380. Bibcode:1940PPS....52..390D. doi:10.1088/0959-5309/52/3/310.
  15. "Color in business, science, and industry. By Deane B. Judd. John Wiley & Sons, Inc., New York, 1952. 401 pp". Journal of the American Pharmaceutical Association (Scientific Ed.). 42 (12): 757. 1953-12-01. doi:10.1002/jps.3030421221. ISSN   0095-9553.
  16. Schuster, Aurhur (1905). "Radiation through a foggy atmosphere". Astrophysical Journal. 21 (1): 1–22. Bibcode:1905ApJ....21....1S. doi: 10.1086/141186 .
  17. Kortüm, Gustav (1969). Reflectance spectroscopy Principles, methods, applications. Berlin: Springer. ISBN   9783642880711. OCLC   714802320.
  18. 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.
  19. 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.
  20. Alcaraz de la Osa, R.; Iparragirre, I.; Ortiz, D.; Saiz, J. M. (March 2020). "The extended Kubelka–Munk theory and its application to spectroscopy". ChemTexts. 6 (1): 2. doi:10.1007/s40828-019-0097-0. S2CID   212642389.
  21. Handbook of near-infrared analysis. Donald A. Burns, Emil W. Ciurczak (2nd ed., rev. and expanded ed.). New York: M. Dekker. 2001. ISBN   0-8247-0534-3. OCLC   46829414.{{cite book}}: CS1 maint: others (link)
  22. Pracht, G.; Weckler, B.; Lutz, H. D. (2003-10-01). "Diffuse Reflection Infrared Fourier Transform (DRIFT) Spectra and High-Temperature DRIFT Spectra of β—Ni(IO3)2·4H2O, Ni(IO3)2·2H2O, Mg(OH)2, and Zn(OH)F: Novel Embedding Materials". Applied Spectroscopy. 57 (10): 1254–1259. Bibcode:2003ApSpe..57.1254P. doi:10.1366/000370203769699126. ISSN   0003-7028. PMID   14639754. S2CID   42585664.
  23. Burger, T.; Kuhn, J.; Caps, R.; Fricke, J. (1997-03-01). "Quantitative Determination of the Scattering and Absorption Coefficients from Diffuse Reflectance and Transmittance Measurements: Application to Pharmaceutical Powders". Applied Spectroscopy. 51 (3): 309–317. Bibcode:1997ApSpe..51..309B. doi:10.1366/0003702971940404. ISSN   0003-7028. S2CID   98159628.
  24. Dahm, Donald J.; Dahm, Kevin D. (2003-12-01). "Illustration of Failure of Continuum Models of Diffuse Reflectance". Journal of Near Infrared Spectroscopy. 11 (6): 479–485. doi:10.1255/jnirs.398. ISSN   0967-0335. S2CID   93926306.
  25. Benford, Frank (1946-09-01). "Radiation in a Diffusing Medium". JOSA. 36 (9): 524–554. doi:10.1364/JOSA.36.000524. PMID   21002043.
  26. Sochorová, Šárka; Jamriška, Ondřej (December 2021). "Practical pigment mixing for digital painting". ACM Transactions on Graphics. 40 (6): 1–11. doi:10.1145/3478513.3480549. S2CID   245013978.