Jump to content

Representative layer theory: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
DJDahm (talk | contribs)
Rewrote article, added a reference
DJDahm (talk | contribs)
Added a History Section
Line 5: Line 5:


Representative layer theory can be used to determine the spectroscopic properties of an assembly of particles from those of the individual particles in the assembly.<ref name=":2">{{Cite journal|last=Dahm|first=Kevin D.|last2=Dahm|first2=Donald J.|date=2004-06-01|title=Relation of Representative Layer Theory to other Theories of Diffuse Reflection|url=https://doi.org/10.1255/jnirs.426|journal=Journal of Near Infrared Spectroscopy|language=en|volume=12|issue=3|pages=189–198|doi=10.1255/jnirs.426|issn=0967-0335}}</ref> The sample is modeled as a series of layers, each of which is parallel to each other and perpendicular to the incident beam. The mathematics of plane parallel layers is then used to extract the desired properties from the data, most notably that of the linear absorption coefficient which behaves in the manner of the coefficient in Beer’s law. The representative layer theory gives a way of performing the calculations for new sample properties by changing the properties of a single layer of the particles, which doesn’t require reworking the mathematics for a sample as a whole.
Representative layer theory can be used to determine the spectroscopic properties of an assembly of particles from those of the individual particles in the assembly.<ref name=":2">{{Cite journal|last=Dahm|first=Kevin D.|last2=Dahm|first2=Donald J.|date=2004-06-01|title=Relation of Representative Layer Theory to other Theories of Diffuse Reflection|url=https://doi.org/10.1255/jnirs.426|journal=Journal of Near Infrared Spectroscopy|language=en|volume=12|issue=3|pages=189–198|doi=10.1255/jnirs.426|issn=0967-0335}}</ref> The sample is modeled as a series of layers, each of which is parallel to each other and perpendicular to the incident beam. The mathematics of plane parallel layers is then used to extract the desired properties from the data, most notably that of the linear absorption coefficient which behaves in the manner of the coefficient in Beer’s law. The representative layer theory gives a way of performing the calculations for new sample properties by changing the properties of a single layer of the particles, which doesn’t require reworking the mathematics for a sample as a whole.

== History ==
The first attempt to account for transmission and reflection of a layered material was carried out by Stokes in about 1860 <ref>{{Cite journal|date=1862-12-31|title=IV. On the intensity of the light reflected from or transmitted through a pile of plates|url=https://royalsocietypublishing.org/doi/10.1098/rspl.1860.0119|journal=Proceedings of the Royal Society of London|language=en|volume=11|pages=545–556|doi=10.1098/rspl.1860.0119|issn=0370-1662}}</ref> and led to some very useful relationships. Lord Rayleigh<ref>{{Cite book|last=Strutt|first=John William|url=http://ebooks.cambridge.org/ref/id/CBO9780511703966|title=Scientific Papers|date=2009|publisher=Cambridge University Press|isbn=978-0-511-70396-6|location=Cambridge|doi=10.1017/cbo9780511703966.009}}</ref> and Mie<ref>{{Cite journal|last=Mie|first=Gustav|date=1908|title=Beiträge zur Optik trüber Medien, speziell kolloidaler Metallösungen|url=https://onlinelibrary.wiley.com/doi/abs/10.1002/andp.19083300302|journal=Annalen der Physik|language=en|volume=330|issue=3|pages=377–445|doi=10.1002/andp.19083300302|issn=1521-3889}}</ref> developed the theory of single scatter to a high degree, but Schuster<ref>{{Cite journal|last=Schuster|first=Arthur|date=1905-01-XX|title=Radiation Through a Foggy Atmosphere|url=http://adsabs.harvard.edu/doi/10.1086/141186|journal=The Astrophysical Journal|language=en|volume=21|pages=1|doi=10.1086/141186|issn=0004-637X}}</ref> was the first to consider multiple scatter. He was concerned with the cloudy atmospheres of stars, and developed a plane-parallel layer model in which the radiation field was divided into forward and backward components. This same model was used much later by [[Kubelka-Munk theory |Kubelka and Munk]], whose names are usually attached to it by spectroscopists.

Following WWII, the field of [[Diffuse reflectance spectroscopy|Reflectance Spectroscopy]] was heavily researched, both theoretically and experimentally.  The remission function, <math>F(R_\infty)</math>, following Kubelka-Munk theory, was the leading contender as the metric of absorption analogous to the absorbance function in transmission absorption spectroscopy. 

The form of the K-M solution originally was: <math>F(R_\infty)\equiv \frac{(1-R_\infty)^2}{2R_\infty} = \frac {a_0}{r_0} </math>, but it was rewritten in terms of linear coefficients by some authors, becoming <math> F(R_\infty)\equiv\frac{(1-R_\infty)^2}{2R_\infty} = \frac {k}{s}</math> , taking <math>k</math> and <math>s</math> as being equivalent to the linear absorption and scattering coefficients as they appear in the Bouguer-Lambert law, even though sources who derived the equations preferred the symbolism <math> \frac {K}S</math> and usually emphasized that <math>K=2k</math> and <math>S</math> was a remission or backscattering parameter, which for the case of diffuse scatter should properly be taken as an integral.<ref>{{Cite book|last=Kortüm|first=Gustav|url=https://www.worldcat.org/oclc/714802320|title=Reflectance spectroscopy Principles, methods, applications.|date=1969|publisher=Springer|isbn=978-3-642-88071-1|location=Berlin,|oclc=714802320}}</ref>

In 1966, in a book entitled Reflectance Spectroscopy, Harry Hecht had pointed out that the formulation <math> F(R_\infty) = \frac {k}{s}</math> led to <math>\log F(R_\infty ) = log(k) - log (s)</math>, which enabled plotting <math>F(R_\infty)</math> "against the wavelength or wavenumber for a particular sample" giving a curve corresponding "to the real absorption determined by transmission measurements, except for a displacement by <math>- log(s)</math> in the ordinate direction." However, in data presented, "the marked deviation in the remission function ... in the region of large extinction is obvious." He listed various reasons given by other authors for this "failure ... to remain valid in strongly absorbing materials", including: "incomplete diffusion in the scattering process"; failure to use "diffuse illumination; "increased proportion of regular reflection"; but concluded that "notwithstanding the above mentioned difficulties, ... the remission function should be a linear function of the concentration at a given wavelength for a constant particle size" though stating that "this discussion has been restricted entirely to the reflectance of homogeneous powder layers" though "equation systems for combination of inhomogeneous layers cannot be solved for the scattering and absorbing properties even in the simple case of a dual combination of sublayers. ... This means that the (Kubelka-Munk) theory fails to include, in an explicit manner, any dependence of reflection on particle size or shaper or refractive index".<ref>{{Cite book|last1=Wendlandt|first1=Wesley Wm |last2=Hechtt|first2=Harry|title=Reflectance Spectroscopy (Chemical Analysis: A Series of Monographs on Analytical Chemistry and Its Applications, Volume 21)|publisher=Interscience Publishers|year=1966|location=New York|pages=72-76}}</ref>

The field of Near infrared spectroscopy (NIR) got its start in 1968, when Karl Norris and co-workers with the Instrumentation Research Lab of the U.S. Department of Agriculture first applied the technology to agricultural products.<ref>{{Cite journal|last=Williams|first=Phil|date=2019-12-XX|title=Karl H. Norris, the Father of Near-Infrared Spectroscopy|url=http://journals.sagepub.com/doi/10.1177/0960336019875883|journal=NIR news|language=en|volume=30|issue=7-8|pages=25–27|doi=10.1177/0960336019875883|issn=0960-3360}}</ref> The USDA discovered how to use NIR empirically, based on available sources, gratings, and detector materials. Even the wavelength range of NIR was empirically set based on the operational range of a PbS detector. Consequently, it was not seen as a rigorous science: it had not evolved in the usual way from research institutions to general usage.<ref>{{Cite book|last=Hindle|first=Peter H|url=https://www.worldcat.org/oclc/1200834251|title=Chap 1: Handbook of near-infrared analysis|date=2021|others=Emil W. Ciurczak, Benoĭt Igne, Jerry Workman, Donald A. Burns|isbn=978-1-351-26988-9|edition=4|location=Boca Raton|oclc=1200834251}}</ref> Even though the Kubelka-Munk theory provided a remission function that could have been used as the absorption metric, Norris selected <math>log(1/R_\infty)</math> for convenience.<ref>{{Cite journal|last=Norris|first=Karl H.|date=2005-12-XX|title=Why Log(1/ R ) for Composition Analysis with Nir?|url=http://journals.sagepub.com/doi/10.1255/nirn.865|journal=NIR news|language=en|volume=16|issue=8|pages=10–13|doi=10.1255/nirn.865|issn=0960-3360}}</ref> He believed that the problem of non-linearity between the metric and concentration was due to particle size (a theoretical concern) and stray light (an instrumental effect).

In 1976, Hecht<ref>{{Cite journal|last=Hecht|first=Harry H|date=1976|title=The Interpretation of Diffuse Reflectance Spectra|journal=JOURNAL OF RESEARCH of he National Bureau of Standards-A. Physics and Chemistry|volume=Vol. 80A, No. 4,|pages=567-583}}</ref> published an exhaustive evaluation of the various theories which were considered to be fairly general. In it, he presented his derivation of the [[Diffuse reflectance spectroscopy#Hecht finite difference formula|Hecht finite difference formula]] by replacing the fundamental differential equations of the Kubelka-Munk theory by the finite difference equations, and obtained: <math>F(R_\infty) = a\biggl( \frac {1}{R} - 1\biggr) - \frac {a^2}{2r}</math>, where <math>a</math> and <math>r</math> are the fractions of light absorbed and remitted by a single layer.  He noted "it is well known that a plot of <math>F(R_\infty)</math>versus <math>K</math> deviates from linearity for high values of <math>K</math>, and it appears that (this equation) can be used to explain the deviations in part", and "does represent an improvement in the range of validity and shows the need to consider the particulate nature of scattering media in developing a more precise theory by which absolute absorptivities can be determined."
== References ==
== References ==
<references />
<references />

Revision as of 19:10, 2 May 2021

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. [1] [2] 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.[3] 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.

Representative layer theory can be used to determine the spectroscopic properties of an assembly of particles from those of the individual particles in the assembly.[4] The sample is modeled as a series of layers, each of which is parallel to each other and perpendicular to the incident beam. The mathematics of plane parallel layers is then used to extract the desired properties from the data, most notably that of the linear absorption coefficient which behaves in the manner of the coefficient in Beer’s law. The representative layer theory gives a way of performing the calculations for new sample properties by changing the properties of a single layer of the particles, which doesn’t require reworking the mathematics for a sample as a whole.

History

The first attempt to account for transmission and reflection of a layered material was carried out by Stokes in about 1860 [5] and led to some very useful relationships. Lord Rayleigh[6] and Mie[7] developed the theory of single scatter to a high degree, but Schuster[8] was the first to consider multiple scatter. He was concerned with the cloudy atmospheres of stars, and developed a plane-parallel layer model in which the radiation field was divided into forward and backward components. This same model was used much later by Kubelka and Munk, whose names are usually attached to it by spectroscopists.

Following WWII, the field of Reflectance Spectroscopy was heavily researched, both theoretically and experimentally.  The remission function, , following Kubelka-Munk theory, was the leading contender as the metric of absorption analogous to the absorbance function in transmission absorption spectroscopy. 

The form of the K-M solution originally was: , but it was rewritten in terms of linear coefficients by some authors, becoming , taking and as being equivalent to the linear absorption and scattering coefficients as they appear in the Bouguer-Lambert law, even though sources who derived the equations preferred the symbolism and usually emphasized that and was a remission or backscattering parameter, which for the case of diffuse scatter should properly be taken as an integral.[9]

In 1966, in a book entitled Reflectance Spectroscopy, Harry Hecht had pointed out that the formulation led to , which enabled plotting "against the wavelength or wavenumber for a particular sample" giving a curve corresponding "to the real absorption determined by transmission measurements, except for a displacement by in the ordinate direction." However, in data presented, "the marked deviation in the remission function ... in the region of large extinction is obvious." He listed various reasons given by other authors for this "failure ... to remain valid in strongly absorbing materials", including: "incomplete diffusion in the scattering process"; failure to use "diffuse illumination; "increased proportion of regular reflection"; but concluded that "notwithstanding the above mentioned difficulties, ... the remission function should be a linear function of the concentration at a given wavelength for a constant particle size" though stating that "this discussion has been restricted entirely to the reflectance of homogeneous powder layers" though "equation systems for combination of inhomogeneous layers cannot be solved for the scattering and absorbing properties even in the simple case of a dual combination of sublayers. ... This means that the (Kubelka-Munk) theory fails to include, in an explicit manner, any dependence of reflection on particle size or shaper or refractive index".[10]

The field of Near infrared spectroscopy (NIR) got its start in 1968, when Karl Norris and co-workers with the Instrumentation Research Lab of the U.S. Department of Agriculture first applied the technology to agricultural products.[11] The USDA discovered how to use NIR empirically, based on available sources, gratings, and detector materials. Even the wavelength range of NIR was empirically set based on the operational range of a PbS detector. Consequently, it was not seen as a rigorous science: it had not evolved in the usual way from research institutions to general usage.[12] Even though the Kubelka-Munk theory provided a remission function that could have been used as the absorption metric, Norris selected for convenience.[13] He believed that the problem of non-linearity between the metric and concentration was due to particle size (a theoretical concern) and stray light (an instrumental effect).

In 1976, Hecht[14] published an exhaustive evaluation of the various theories which were considered to be fairly general. In it, he presented his derivation of the Hecht finite difference formula by replacing the fundamental differential equations of the Kubelka-Munk theory by the finite difference equations, and obtained: , where and are the fractions of light absorbed and remitted by a single layer.  He noted "it is well known that a plot of versus deviates from linearity for high values of , and it appears that (this equation) can be used to explain the deviations in part", and "does represent an improvement in the range of validity and shows the need to consider the particulate nature of scattering media in developing a more precise theory by which absolute absorptivities can be determined."

References

  1. ^ Dahm, Donald J.; Dahm, Kevin D. (1999-06-01). "Representative Layer Theory for Diffuse Reflectance". Applied Spectroscopy. 53 (6): 647–654. doi:10.1366/0003702991947298. ISSN 0003-7028.
  2. ^ Dahm, Donald J.; Dahm, Kevin D.; Norris, Karl H. (2000-06-01). "Test of the Representative Layer Theory of Diffuse Reflectance Using Plane Parallel Samples". Journal of Near Infrared Spectroscopy. 8 (3): 171–181. doi:10.1255/jnirs.276. ISSN 0967-0335.
  3. ^ Bohren, Craig F.; Huffman, Donald R. (1998-04-23). Absorption and Scattering of Light by Small Particles (1 ed.). Wiley. doi:10.1002/9783527618156. ISBN 978-0-471-29340-8.
  4. ^ Dahm, Kevin D.; Dahm, Donald J. (2004-06-01). "Relation of Representative Layer Theory to other Theories of Diffuse Reflection". Journal of Near Infrared Spectroscopy. 12 (3): 189–198. doi:10.1255/jnirs.426. ISSN 0967-0335.
  5. ^ "IV. 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. 1862-12-31. doi:10.1098/rspl.1860.0119. ISSN 0370-1662.
  6. ^ Strutt, John William (2009). Scientific Papers. Cambridge: Cambridge University Press. doi:10.1017/cbo9780511703966.009. ISBN 978-0-511-70396-6.
  7. ^ Mie, Gustav (1908). "Beiträge zur Optik trüber Medien, speziell kolloidaler Metallösungen". Annalen der Physik. 330 (3): 377–445. doi:10.1002/andp.19083300302. ISSN 1521-3889.
  8. ^ Schuster, Arthur (1905-01-XX). "Radiation Through a Foggy Atmosphere". The Astrophysical Journal. 21: 1. doi:10.1086/141186. ISSN 0004-637X. {{cite journal}}: Check date values in: |date= (help)
  9. ^ Kortüm, Gustav (1969). Reflectance spectroscopy Principles, methods, applications. Berlin,: Springer. ISBN 978-3-642-88071-1. OCLC 714802320.{{cite book}}: CS1 maint: extra punctuation (link)
  10. ^ Wendlandt, Wesley Wm; Hechtt, Harry (1966). Reflectance Spectroscopy (Chemical Analysis: A Series of Monographs on Analytical Chemistry and Its Applications, Volume 21). New York: Interscience Publishers. pp. 72–76.
  11. ^ Williams, Phil (2019-12-XX). "Karl H. Norris, the Father of Near-Infrared Spectroscopy". NIR news. 30 (7–8): 25–27. doi:10.1177/0960336019875883. ISSN 0960-3360. {{cite journal}}: Check date values in: |date= (help)
  12. ^ Hindle, Peter H (2021). Chap 1: Handbook of near-infrared analysis. Emil W. Ciurczak, Benoĭt Igne, Jerry Workman, Donald A. Burns (4 ed.). Boca Raton. ISBN 978-1-351-26988-9. OCLC 1200834251.{{cite book}}: CS1 maint: location missing publisher (link)
  13. ^ Norris, Karl H. (2005-12-XX). "Why Log(1/ R ) for Composition Analysis with Nir?". NIR news. 16 (8): 10–13. doi:10.1255/nirn.865. ISSN 0960-3360. {{cite journal}}: Check date values in: |date= (help)
  14. ^ Hecht, Harry H (1976). "The Interpretation of Diffuse Reflectance Spectra". JOURNAL OF RESEARCH of he National Bureau of Standards-A. Physics and Chemistry. Vol. 80A, No. 4, : 567–583. {{cite journal}}: |volume= has extra text (help)CS1 maint: extra punctuation (link)