Helmholtz reciprocity

Last updated

The Helmholtz reciprocity principle describes how a ray of light and its reverse ray encounter matched optical adventures, such as reflections, refractions, and absorptions in a passive medium, or at an interface. It does not apply to moving, non-linear, or magnetic media.

Contents

For example, incoming and outgoing light can be considered as reversals of each other, [1] without affecting the bidirectional reflectance distribution function (BRDF) [2] outcome. If light was measured with a sensor and that light reflected on a material with a BRDF that obeys the Helmholtz reciprocity principle one would be able to swap the sensor and light source and the measurement of flux would remain equal.

In the computer graphics scheme of global illumination, the Helmholtz reciprocity principle is important if the global illumination algorithm reverses light paths (for example raytracing versus classic light path tracing).

Physics

The Stokes–Helmholtz reversion–reciprocity principle [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [1] [14] [15] [16] [17] [18] [19] [20] [21] [22] [ excessive citations ] was stated in part by Stokes (1849) [3] and with reference to polarization on page 169 [4] of Hermann Helmholtz's Handbuch der physiologischen Optik of 1856 as cited by Gustav Kirchhoff [8] and by Max Planck. [13]

As cited by Kirchhoff in 1860, the principle is translated as follows:

A ray of light proceeding from point 1 arrives at point 2 after suffering any number of refractions, reflections, &c. At point 1 let any two perpendicular planes a1, b1 be taken in the direction of the ray; and let the vibrations of the ray be divided into two parts, one in each of these planes. Take similar planes a2, b2 in the ray at point 2; then the following proposition may be demonstrated. If when the quantity of light i polarized in the plane a1 proceeds from 1 in the direction of the given ray, that part k thereof of light polarized in a2 arrives at 2, then, conversely, if the quantity of light i polarized in a2 proceeds from 2, the same quantity of light k polarized in a1 [Kirchhoff's published text here corrected by Wikipedia editor to agree with Helmholtz's 1867 text] will arrive at 1. [8]

Simply put, in suitable conditions, the principle states that the source and observation point may be switched without changing the measured intensity. Intuitively, "If I can see you, you can see me." Like the principles of thermodynamics, in suitable conditions, this principle is reliable enough to use as a check on the correct performance of experiments, in contrast with the usual situation in which the experiments are tests of a proposed law. [1] [12]

In his magisterial proof [23] of the validity of Kirchhoff's law of equality of radiative emissivity and absorptivity, [24] Planck makes repeated and essential use of the Stokes–Helmholtz reciprocity principle. Rayleigh stated the basic idea of reciprocity as a consequence of the linearity of propagation of small vibrations, light consisting of sinusoidal vibrations in a linear medium. [9] [10] [11] [12]

When there are magnetic fields in the path of the ray, the principle does not apply. [4] Departure of the optical medium from linearity also causes departure from Helmholtz reciprocity, as well as the presence of moving objects in the path of the ray.

Helmholtz reciprocity referred originally to light. This is a particular form of electromagnetism that may be called far-field radiation. For this, the electric and magnetic fields do not need distinct descriptions, because they propagate feeding each other evenly. So the Helmholtz principle is a more simply described special case of electromagnetic reciprocity in general, which is described by distinct accounts of the interacting electric and magnetic fields. The Helmholtz principle rests mainly on the linearity and superposability of the light field, and it has close analogues in non-electromagnetic linear propagating fields, such as sound. It was discovered before the electromagnetic nature of light became known. [9] [10] [11] [12]

The Helmholtz reciprocity theorem has been rigorously proven in a number of ways, [25] [26] [27] generally making use of quantum mechanical time-reversal symmetry. As these more mathematically complicated proofs may detract from the simplicity of the theorem, A.P Pogany and P. S. Turner have proven it in only a few steps using a Born series. [28] Assuming a light source at a point A and an observation point O, with various scattering points between them, the Schrödinger equation may be used to represent the resulting wave function in space:

By applying a Green's function, the above equation can be solved for the wave function in an integral (and thus iterative) form:

where

.

Next, it is valid to assume the solution inside the scattering medium at point O may be approximated by a Born series, making use of the Born approximation in scattering theory. In doing so, the series may be iterated through in the usual way to generate the following integral solution:

Noting again the form of the Green's function, it is apparent that switching and in the above form will not change the result; that is to say, , which is the mathematical statement of the reciprocity theorem: switching the light source A and observation point O does not alter the observed wave function.

Applications

One simple yet important implication of this reciprocity principle is that any light directed through a lens in one direction (from object to image plane) is optically equal to its conjugate, i.e. light being directed through the same set-up but in the opposite direction. An electron being focused through any series of optical components does not “care” from which direction it comes; as long as the same optical events happen to it, the resulting wave function will be the same. For that reason, this principle has important applications in the field of transmission electron microscopy (TEM). The notion that conjugate optical processes produce equivalent results allows the microscope user to grasp a deeper understanding of, and have considerable flexibility in, techniques involving electron diffraction, Kikuchi patterns, [29] dark-field images, [28] and others.

An important caveat to note is that in a situation where electrons lose energy after interacting with the scattering medium of the sample, there is not time-reversal symmetry. Therefore, reciprocity only truly applies in situations of elastic scattering. In the case of inelastic scattering with small energy loss, it can be shown that reciprocity may be used to approximate intensity (rather than wave amplitude). [28] So in very thick samples or samples in which inelastic scattering dominates, the benefits of using reciprocity for the previously mentioned TEM applications are no longer valid. Furthermore, it has been demonstrated experimentally that reciprocity does apply in a TEM under the right conditions, [28] but the underlying physics of the principle dictates that reciprocity can only be truly exact if ray transmission occurs through only scalar fields, i.e. no magnetic fields. We can therefore conclude that the distortions to reciprocity due to magnetic fields of the electromagnetic lenses in TEM may be ignored under typical operating conditions. [30] However, users should be careful not to apply reciprocity to magnetic imaging techniques, TEM of ferromagnetic materials, or extraneous TEM situations without careful consideration. Generally, polepieces for TEM are designed using finite element analysis of generated magnetic fields to ensure symmetry.  

Magnetic objective lens systems have been used in TEM to achieve atomic-scale resolution while maintaining a magnetic field free environment at the plane of the sample, [31] but the method of doing so still requires a large magnetic field above (and below) the sample, thus negating any reciprocity enhancement effects that one might expect. This system works by placing the sample in between the front and back objective lens polepieces, as in an ordinary TEM, but the two polepieces are kept in exact mirror symmetry with respect to the sample plane between them. Meanwhile, their excitation polarities are exactly opposite, generating magnetic fields that cancel almost perfectly at the plane of the sample. However, since they do not cancel elsewhere, the electron trajectory must still pass through magnetic fields.

Reciprocity can also be used to understand the main difference between TEM and scanning transmission electron microscopy (STEM), which is characterized in principle by switching the position of the electron source and observation point. This is effectively the same as reversing time on a TEM so that electrons travel in the opposite direction. Therefore, under appropriate conditions (in which reciprocity does apply), knowledge of TEM imaging can be useful in taking and interpreting images with STEM.

See also

Related Research Articles

<span class="mw-page-title-main">Diffraction</span> Phenomenon of the motion of waves

Diffraction is the interference or bending of waves around the corners of an obstacle or through an aperture into the region of geometrical shadow of the obstacle/aperture. The diffracting object or aperture effectively becomes a secondary source of the propagating wave. Italian scientist Francesco Maria Grimaldi coined the word diffraction and was the first to record accurate observations of the phenomenon in 1660.

<span class="mw-page-title-main">Huygens–Fresnel principle</span> Method of analysis

The Huygens–Fresnel principle states that every point on a wavefront is itself the source of spherical wavelets, and the secondary wavelets emanating from different points mutually interfere. The sum of these spherical wavelets forms a new wavefront. As such, the Huygens-Fresnel principle is a method of analysis applied to problems of luminous wave propagation both in the far-field limit and in near-field diffraction as well as reflection.

<span class="mw-page-title-main">Quantum electrodynamics</span> Quantum field theory of electromagnetism

In particle physics, quantum electrodynamics (QED) is the relativistic quantum field theory of electrodynamics. In essence, it describes how light and matter interact and is the first theory where full agreement between quantum mechanics and special relativity is achieved. QED mathematically describes all phenomena involving electrically charged particles interacting by means of exchange of photons and represents the quantum counterpart of classical electromagnetism giving a complete account of matter and light interaction.

In electrodynamics, elliptical polarization is the polarization of electromagnetic radiation such that the tip of the electric field vector describes an ellipse in any fixed plane intersecting, and normal to, the direction of propagation. An elliptically polarized wave may be resolved into two linearly polarized waves in phase quadrature, with their polarization planes at right angles to each other. Since the electric field can rotate clockwise or counterclockwise as it propagates, elliptically polarized waves exhibit chirality.

<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.

Electron density or electronic density is the measure of the probability of an electron being present at an infinitesimal element of space surrounding any given point. It is a scalar quantity depending upon three spatial variables and is typically denoted as either or . The density is determined, through definition, by the normalised Failed to parse : N -electron wavefunction which itself depends upon variables. Conversely, the density determines the wave function modulo up to a phase factor, providing the formal foundation of density functional theory.

Fourier optics is the study of classical optics using Fourier transforms (FTs), in which the waveform being considered is regarded as made up of a combination, or superposition, of plane waves. It has some parallels to the Huygens–Fresnel principle, in which the wavefront is regarded as being made up of a combination of spherical wavefronts whose sum is the wavefront being studied. A key difference is that Fourier optics considers the plane waves to be natural modes of the propagation medium, as opposed to Huygens–Fresnel, where the spherical waves originate in the physical medium.

<span class="mw-page-title-main">Magnetic moment</span> Magnetic strength and orientation of an object that produces a magnetic field

In electromagnetism, the magnetic moment is the magnetic strength and orientation of a magnet or other object that produces a magnetic field. Examples of objects that have magnetic moments include loops of electric current, permanent magnets, elementary particles, composite particles, various molecules, and many astronomical objects.

<span class="mw-page-title-main">Fine structure</span> Details in the emission spectrum of an atom

In atomic physics, the fine structure describes the splitting of the spectral lines of atoms due to electron spin and relativistic corrections to the non-relativistic Schrödinger equation. It was first measured precisely for the hydrogen atom by Albert A. Michelson and Edward W. Morley in 1887, laying the basis for the theoretical treatment by Arnold Sommerfeld, introducing the fine-structure constant.

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

The Compton wavelength is a quantum mechanical property of a particle, defined as the wavelength of a photon whose energy is the same as the rest energy of that particle. It was introduced by Arthur Compton in 1923 in his explanation of the scattering of photons by electrons.

<span class="mw-page-title-main">Gauge fixing</span> Procedure of coping with redundant degrees of freedom in physical field theories

In the physics of gauge theories, gauge fixing denotes a mathematical procedure for coping with redundant degrees of freedom in field variables. By definition, a gauge theory represents each physically distinct configuration of the system as an equivalence class of detailed local field configurations. Any two detailed configurations in the same equivalence class are related by a gauge transformation, equivalent to a shear along unphysical axes in configuration space. Most of the quantitative physical predictions of a gauge theory can only be obtained under a coherent prescription for suppressing or ignoring these unphysical degrees of freedom.

In classical electromagnetism, reciprocity refers to a variety of related theorems involving the interchange of time-harmonic electric current densities (sources) and the resulting electromagnetic fields in Maxwell's equations for time-invariant linear media under certain constraints. Reciprocity is closely related to the concept of symmetric operators from linear algebra, applied to electromagnetism.

Generally in scattering theory and in particular in quantum mechanics, the Born approximation consists of taking the incident field in place of the total field as the driving field at each point in the scatterer. The Born approximation is named after Max Born who proposed this approximation in early days of quantum theory development.

In physics, the optical theorem is a general law of wave scattering theory, which relates the forward scattering amplitude to the total cross section of the scatterer. It is usually written in the form

<span class="mw-page-title-main">Hofstadter's butterfly</span> Fractal describing the theorised behaviour of electrons in a magnetic field

In condensed matter physics, Hofstadter's butterfly is a graph of the spectral properties of non-interacting two-dimensional electrons in a perpendicular magnetic field in a lattice. The fractal, self-similar nature of the spectrum was discovered in the 1976 Ph.D. work of Douglas Hofstadter and is one of the early examples of modern scientific data visualization. The name reflects the fact that, as Hofstadter wrote, "the large gaps [in the graph] form a very striking pattern somewhat resembling a butterfly."

The theoretical and experimental justification for the Schrödinger equation motivates the discovery of the Schrödinger equation, the equation that describes the dynamics of nonrelativistic particles. The motivation uses photons, which are relativistic particles with dynamics described by Maxwell's equations, as an analogue for all types of particles.

In quantum mechanics, the Pauli equation or Schrödinger–Pauli equation is the formulation of the Schrödinger equation for spin-½ particles, which takes into account the interaction of the particle's spin with an external electromagnetic field. It is the non-relativistic limit of the Dirac equation and can be used where particles are moving at speeds much less than the speed of light, so that relativistic effects can be neglected. It was formulated by Wolfgang Pauli in 1927.

In physics, relativistic quantum mechanics (RQM) is any Poincaré covariant formulation of quantum mechanics (QM). This theory is applicable to massive particles propagating at all velocities up to those comparable to the speed of light c, and can accommodate massless particles. The theory has application in high energy physics, particle physics and accelerator physics, as well as atomic physics, chemistry and condensed matter physics. Non-relativistic quantum mechanics refers to the mathematical formulation of quantum mechanics applied in the context of Galilean relativity, more specifically quantizing the equations of classical mechanics by replacing dynamical variables by operators. Relativistic quantum mechanics (RQM) is quantum mechanics applied with special relativity. Although the earlier formulations, like the Schrödinger picture and Heisenberg picture were originally formulated in a non-relativistic background, a few of them also work with special relativity.

The multislice algorithm is a method for the simulation of the elastic interaction of an electron beam with matter, including all multiple scattering effects. The method is reviewed in the book by Cowley, and also the work by Ishizuka. The algorithm is used in the simulation of high resolution transmission electron microscopy micrographs, and serves as a useful tool for analyzing experimental images. This article describes some relevant background information, the theoretical basis of the technique, approximations used, and several software packages that implement this technique. Some of the advantages and limitations of the technique and important considerations that need to be taken into account are described.

References

  1. 1 2 3 Hapke, B. (1993). Theory of Reflectance and Emittance Spectroscopy, Cambridge University Press, Cambridge UK, ISBN   0-521-30789-9, Section 10C, pages 263-264.
  2. Hapke, B. (1993). Theory of Reflectance and Emittance Spectroscopy, Cambridge University Press, Cambridge UK, ISBN   0-521-30789-9, Chapters 8-9, pages 181-260.
  3. 1 2 Stokes, G.G. (1849). "On the perfect blackness of the central spot in Newton's rings, and on the verification of Fresnel's formulae for the intensities of reflected and refracted rays". Cambridge and Dublin Mathematical Journal. new series. 4: 1-14.
  4. 1 2 3 Helmholtz, H. von (1856). Handbuch der physiologischen Optik, first edition cited by Planck, Leopold Voss, Leipzig, volume 1, page 169.
  5. Helmholtz, H. von (1903). Vorlesungen über Theorie der Wärme, edited by F. Richarz, Johann Ambrosius Barth, Leipzig, pages 158-162.
  6. Helmholtz, H. (1859/60). Theorie der Luftschwingungen in Röhren mit offenen Enden, Crelle's Journal für die reine und angewandte Mathematik57(1): 1-72, page 29.
  7. Stewart, B. (1858). An account of some experiments on radiant heat, involving an extension of Professor Prevost's theory of exchanges, Trans. Roy. Soc. Edinburgh22 (1): 1-20, page 18.
  8. 1 2 3 Kirchhoff, G. (1860). On the Relation between the Radiating and Absorbing Powers of different Bodies for Light and Heat, Ann. Phys., 119: 275-301, at page 287 , translated by F. Guthrie, Phil. Mag. Series 4, 20:2-21, at page 9.
  9. 1 2 3 Strutt, J.W. (Lord Rayleigh) (1873). Some general theorems relating to vibrations, Proc. Lond. Math. Soc.4: 357-368, pages 366-368.
  10. 1 2 3 Rayleigh, Lord (1876). On the application of the Principle of Reciprocity to acoustics, Proc. Roy. Soc. A, 25: 118-122.
  11. 1 2 3 Strutt, J.W., Baron Rayleigh (1894/1945). The Theory of Sound, second revised edition, Dover, New York, volume 1, sections 107-111a.
  12. 1 2 3 4 Rayleigh, Lord (1900). On the law of reciprocity in diffuse reflection, Phil. Mag. series 5, 49: 324-325.
  13. 1 2 Planck, M. (1914). The Theory of Heat Radiation, second edition translated by M. Masius, P. Blakiston's Son and Co., Philadelphia, page 35.
  14. Minnaert, M. (1941). The reciprocity principle in lunar photometry, Astrophysical Journal93: 403-410.
  15. Mahan, A.I. (1943). A mathematical proof of Stokes' reversibility principle, J. Opt. Soc. Am., 33(11): 621-626.
  16. Chandrasekhar, S. (1950). Radiative Transfer, Oxford University Press, Oxford, pages 20-21, 171-177, 182.
  17. Tingwaldt, C.P. (1952). Über das Helmholtzsche Reziprozitätsgesetz in der Optik, Optik, 9(6): 248-253.
  18. Levi, L. (1968). Applied Optics: A Guide to Optical System Design, 2 volumes, Wiley, New York, volume 1, page 84.
  19. Clarke, F.J.J., Parry, D.J. (1985). Helmholtz reciprocity: its validity and application to reflectometry, Lighting Research & Technology , 17(1): 1-11.
  20. Lekner, J. (1987). Theory of reflection, Martinus Nijhoff, Dordrecht, ISBN   90-247-3418-5, pages 33-37.
  21. Born, M., Wolf, E. (1999). Principles of Optics: Electromagnetic theory of propagation, interference and diffraction of light, 7th edition, Cambridge University Press, ISBN   0-521-64222-1, page 423.
  22. Potton, R J (April 27, 2004). "Reciprocity in optics". Reports on Progress in Physics. IOP Publishing. 67 (5): 717–754. Bibcode:2004RPPh...67..717P. doi:10.1088/0034-4885/67/5/r03. ISSN   0034-4885. S2CID   250849465.
  23. Planck, M. (1914). The Theory of Heat Radiation, second edition translated by M. Masius, P. Blakiston's Son and Co., Philadelphia, pages 35, 38,39.
  24. Kirchhoff, G. (1860). On the Relation between the Radiating and Absorbing Powers of different Bodies for Light and Heat, Ann. Phys., 119: 275-301 , translated by F. Guthrie, Phil. Mag. Series 4, 20:2-21.
  25. Helmholtz, Hermann von (1867). a, Hermann von Helmholtz u (ed.). Handbuch der physiologischen Optik (in German). Leipzig: L. Voss.
  26. Wells, Oliver C. (July 23, 2008). "Reciprocity between the reflection electron microscope and the low‐loss scanning electron microscope". Applied Physics Letters. 37 (6): 507–510. doi:10.1063/1.91992. ISSN   0003-6951.
  27. Spindler, Paul (de Chemnitz) Auteur du texte; Meyer, Georg (1857-1950) Auteur du texte; Meerburg, Jacob Hendrik Auteur du texte (1860). "Annalen der Physik". Gallica. Retrieved December 11, 2019.
  28. 1 2 3 4 Pogany, A. P.; Turner, P. S. (January 23, 1968). "Reciprocity in electron diffraction and microscopy". Acta Crystallographica Section A. 24 (1): 103–109. Bibcode:1968AcCrA..24..103P. doi:10.1107/S0567739468000136. ISSN   1600-5724.
  29. Kainuma, Y. (May 10, 1955). "The Theory of Kikuchi patterns". Acta Crystallographica. 8 (5): 247–257. doi: 10.1107/S0365110X55000832 . ISSN   0365-110X.
  30. Hren, John J; Goldstein, Joseph I; Joy, David C, eds. (1979). Introduction to Analytical Electron Microscopy | SpringerLink (PDF). doi:10.1007/978-1-4757-5581-7. ISBN   978-1-4757-5583-1.
  31. Shibata, N.; Kohno, Y.; Nakamura, A.; Morishita, S.; Seki, T.; Kumamoto, A.; Sawada, H.; Matsumoto, T.; Findlay, S. D.; Ikuhara, Y. (May 24, 2019). "Atomic resolution electron microscopy in a magnetic field free environment". Nature Communications. 10 (1): 2308. Bibcode:2019NatCo..10.2308S. doi:10.1038/s41467-019-10281-2. ISSN   2041-1723. PMC   6534592 . PMID   31127111.