Reciprocity (electromagnetism)

Last updated

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.

Contents

Perhaps the most common and general such theorem is Lorentz reciprocity (and its various special cases such as Rayleigh-Carson reciprocity), named after work by Hendrik Lorentz in 1896 following analogous results regarding sound by Lord Rayleigh and light by Helmholtz (Potton, 2004). Loosely, it states that the relationship between an oscillating current and the resulting electric field is unchanged if one interchanges the points where the current is placed and where the field is measured. For the specific case of an electrical network, it is sometimes phrased as the statement that voltages and currents at different points in the network can be interchanged. More technically, it follows that the mutual impedance of a first circuit due to a second is the same as the mutual impedance of the second circuit due to the first.

Reciprocity is useful in optics, which (apart from quantum effects) can be expressed in terms of classical electromagnetism, but also in terms of radiometry.

There is also an analogous theorem in electrostatics, known as Green's reciprocity, relating the interchange of electric potential and electric charge density.

Forms of the reciprocity theorems are used in many electromagnetic applications, such as analyzing electrical networks and antenna systems. [1] For example, reciprocity implies that antennas work equally well as transmitters or receivers, and specifically that an antenna's radiation and receiving patterns are identical. Reciprocity is also a basic lemma that is used to prove other theorems about electromagnetic systems, such as the symmetry of the impedance matrix and scattering matrix, symmetries of Green's functions for use in boundary-element and transfer-matrix computational methods, as well as orthogonality properties of harmonic modes in waveguide systems (as an alternative to proving those properties directly from the symmetries of the eigen-operators).

Lorentz reciprocity

Specifically, suppose that one has a current density that produces an electric field and a magnetic field where all three are periodic functions of time with angular frequency ω, and in particular they have time-dependence Suppose that we similarly have a second current at the same frequency ω which (by itself) produces fields and The Lorentz reciprocity theorem then states, under certain simple conditions on the materials of the medium described below, that for an arbitrary surface S enclosing a volume V:

Equivalently, in differential form (by the divergence theorem):

This general form is commonly simplified for a number of special cases. In particular, one usually assumes that and are localized (i.e. have compact support), and that there are no incoming waves from infinitely far away. In this case, if one integrates throughout space then the surface-integral terms cancel (see below) and one obtains:

This result (along with the following simplifications) is sometimes called the Rayleigh-Carson reciprocity theorem, after Lord Rayleigh's work on sound waves and an extension by Carson (1924; 1930) to applications for radio frequency antennas. Often, one further simplifies this relation by considering point-like dipole sources, in which case the integrals disappear and one simply has the product of the electric field with the corresponding dipole moments of the currents. Or, for wires of negligible thickness, one obtains the applied current in one wire multiplied by the resulting voltage across another and vice versa; see also below.

Another special case of the Lorentz reciprocity theorem applies when the volume V entirely contains both of the localized sources (or alternatively if V intersects neither of the sources). In this case:

In practical problems, there are another more generalized forms of Lorentz and other reciprocity relations, in which, in addition to electric current density , magnetic current density is also used. These types of reciprocity relations are usually discussed in electrical engineering literature. [2] [3] [4] [5] [6]

Reciprocity for electrical networks

Above, Lorentz reciprocity was phrased in terms of an externally applied current source and the resulting field. Often, especially for electrical networks, one instead prefers to think of an externally applied voltage and the resulting currents. The Lorentz reciprocity theorem describes this case as well, assuming ohmic materials (i.e. currents that respond linearly to the applied field) with a 3×3 conductivity matrix σ that is required to be symmetric, which is implied by the other conditions below. In order to properly describe this situation, one must carefully distinguish between the externally applied fields (from the driving voltages) and the total fields that result (King, 1963).

More specifically, the above only consisted of external "source" terms introduced into Maxwell's equations. We now denote this by to distinguish it from the total current produced by both the external source and by the resulting electric fields in the materials. If this external current is in a material with a conductivity σ, then it corresponds to an externally applied electric field where, by definition of σ:

Moreover, the electric field above only consisted of the response to this current, and did not include the "external" field Therefore, we now denote the field from before as where the total field is given by

Now, the equation on the left-hand side of the Lorentz reciprocity theorem can be rewritten by moving the σ from the external current term to the response field terms and also adding and subtracting a term, to obtain the external field multiplied by the total current

For the limit of thin wires, this gives the product of the externally applied voltage (1) multiplied by the resulting total current (2) and vice versa. In particular, the Rayleigh-Carson reciprocity theorem becomes a simple summation:

where and I denote the complex amplitudes of the AC applied voltages and the resulting currents, respectively, in a set of circuit elements (indexed by n) for two possible sets of voltages and

Most commonly, this is simplified further to the case where each system has a single voltage source at and Then the theorem becomes simply

or in words:

The current at position (1) from a voltage at (2) is identical to the current at (2) from the same voltage at (1).

Conditions and proof of Lorentz reciprocity

The Lorentz reciprocity theorem is simply a reflection of the fact that the linear operator relating and at a fixed frequency (in linear media):

where

is usually a symmetric operator under the "inner product" for vector fields and [7] (Technically, this unconjugated form is not a true inner product because it is not real-valued for complex-valued fields, but that is not a problem here. In this sense, the operator is not truly Hermitian but is rather complex-symmetric.) This is true whenever the permittivity ε and the magnetic permeability μ, at the given ω, are symmetric 3×3 matrices (symmetric rank-2 tensors) – this includes the common case where they are scalars (for isotropic media), of course. They need not be real – complex values correspond to materials with losses, such as conductors with finite conductivity σ (which is included in ε via ) – and because of this, the reciprocity theorem does not require time reversal invariance. The condition of symmetric ε and μ matrices is almost always satisfied; see below for an exception.

For any Hermitian operator under an inner product , we have by definition, and the Rayleigh-Carson reciprocity theorem is merely the vectorial version of this statement for this particular operator that is, The Hermitian property of the operator here can be derived by integration by parts. For a finite integration volume, the surface terms from this integration by parts yield the more-general surface-integral theorem above. In particular, the key fact is that, for vector fields and integration by parts (or the divergence theorem) over a volume V enclosed by a surface S gives the identity:

This identity is then applied twice to to yield plus the surface term, giving the Lorentz reciprocity relation.

Conditions and proof of Lorenz reciprocity using Maxwell's equations and vector operations [8]

We shall prove a general form of the electromagnetic reciprocity theorem due to Lorenz which states that fields and generated by two different sinusoidal current densities respectively and of the same frequency, satisfy the condition

Let us take a region in which dielectric constant and permeability may be functions of position but not of time. Maxwell's equations, written in terms of the total fields, currents and charges of the region describe the electromagnetic behavior of the region. The two curl equations are:

Under steady constant frequency conditions we get from the two curl equations the Maxwell's equations for the Time-Periodic case:

It must be recognized that the symbols in the equations of this article represent the complex multipliers of , giving the in-phase and out-of-phase parts with respect to the chosen reference. The complex vector multipliers of may be called vector phasors by analogy to the complex scalar quantities which are commonly referred to as phasors.

An equivalence of vector operations shows that

for every vectors and

If we apply this equivalence to and we get:

If products in the Time-Periodic equations are taken as indicated by this last equivalence, and added,

This now may be integrated over the volume of concern,

From the divergence theorem the volume integral of equals the surface integral of over the boundary.

This form is valid for general media, but in the common case of linear, isotropic, time-invariant materials, ε is a scalar independent of time. Then generally as physical magnitudes and

Last equation then becomes

In an exactly analogous way we get for vectors and the following expression:

Subtracting the two last equations by members we get

and equivalently in differential form

Q.E.D.

Surface-term cancellation

The cancellation of the surface terms on the right-hand side of the Lorentz reciprocity theorem, for an integration over all space, is not entirely obvious but can be derived in a number of ways. A rigorous treatment of the surface integral takes into account the causality of interacting wave field states: The surface-integral contribution at infinity vanishes for the time-convolution interaction of two causal wave fields only (the time-correlation interaction leads to a non-zero contribution). [9]

Another simple argument would be that the fields goes to zero at infinity for a localized source, but this argument fails in the case of lossless media: in the absence of absorption, radiated fields decay inversely with distance, but the surface area of the integral increases with the square of distance, so the two rates balance one another in the integral.

Instead, it is common (e.g. King, 1963) to assume that the medium is homogeneous and isotropic sufficiently far away. In this case, the radiated field asymptotically takes the form of planewaves propagating radially outward (in the direction) with and where Z is the scalar impedance of the surrounding medium. Then it follows that which by a simple vector identity equals Similarly, and the two terms cancel one another.

The above argument shows explicitly why the surface terms can cancel, but lacks generality. Alternatively, one can treat the case of lossless surrounding media with radiation boundary conditions imposed via the limiting absorption principle (LAP): Taking the limit as the losses (the imaginary part of ε) go to zero. For any nonzero loss, the fields decay exponentially with distance and the surface integral vanishes, regardless of whether the medium is homogeneous. Since the left-hand side of the Lorentz reciprocity theorem vanishes for integration over all space with any non-zero losses, it must also vanish in the limit as the losses go to zero. (Note that the LAP implicitly imposes the Sommerfeld radiation condition of zero incoming waves from infinity, because otherwise even an arbitrarily small loss would eliminate the incoming waves and the limit would not give the lossless solution.)

Reciprocity and Green's function

The inverse of the operator i.e., in (which requires a specification of the boundary conditions at infinity in a lossless system), has the same symmetry as and is essentially a Green's function convolution. So, another perspective on Lorentz reciprocity is that it reflects the fact that convolution with the electromagnetic Green's function is a complex-symmetric (or anti-Hermitian, below) linear operation under the appropriate conditions on ε and μ. More specifically, the Green's function can be written as giving the n-th component of at from a point dipole current in the m-th direction at (essentially, gives the matrix elements of ), and Rayleigh-Carson reciprocity is equivalent to the statement that Unlike it is not generally possible to give an explicit formula for the Green's function (except in special cases such as homogeneous media), but it is routinely computed by numerical methods.

Lossless magneto-optic materials

One case in which ε is not a symmetric matrix is for magneto-optic materials, in which case the usual statement of Lorentz reciprocity does not hold (see below for a generalization, however). If we allow magneto-optic materials, but restrict ourselves to the situation where material absorption is negligible, then ε and μ are in general 3×3 complex Hermitian matrices. In this case, the operator is Hermitian under the conjugated inner product and a variant of the reciprocity theorem[ citation needed ] still holds:

where the sign changes come from the in the equation above, which makes the operator anti-Hermitian (neglecting surface terms). For the special case of this gives a re-statement of conservation of energy or Poynting's theorem (since here we have assumed lossless materials, unlike above): The time-average rate of work done by the current (given by the real part of ) is equal to the time-average outward flux of power (the integral of the Poynting vector). By the same token, however, the surface terms do not in general vanish if one integrates over all space for this reciprocity variant, so a Rayleigh-Carson form does not hold without additional assumptions.

The fact that magneto-optic materials break Rayleigh-Carson reciprocity is the key to devices such as Faraday isolators and circulators. A current on one side of a Faraday isolator produces a field on the other side but not vice versa.

Generalization to non-symmetric materials

For a combination of lossy and magneto-optic materials, and in general when the ε and μ tensors are neither symmetric nor Hermitian matrices, one can still obtain a generalized version of Lorentz reciprocity by considering and to exist in different systems.

In particular, if satisfy Maxwell's equations at ω for a system with materials and satisfy Maxwell's equations at ω for a system with materials where denotes the transpose, then the equation of Lorentz reciprocity holds. This can be further generalized to bi-anisotropic materials by transposing the full 6×6 susceptibility tensor. [10]

Exceptions to reciprocity

For nonlinear media, no reciprocity theorem generally holds. Reciprocity also does not generally apply for time-varying ("active") media; for example, when ε is modulated in time by some external process. (In both of these cases, the frequency ω is not generally a conserved quantity.)

Feld-Tai reciprocity

In 1992, a closely related reciprocity theorem was articulated independently by Y.A. Feld [11] and C.T. Tai, [12] and is known as Feld-Tai reciprocity or the Feld-Tai lemma. t relates two time-harmonic localized current sources and the resulting magnetic fields:

However, the Feld-Tai lemma is only valid under much more restrictive conditions than Lorentz reciprocity. It generally requires time-invariant linear media with an isotropic homogeneous impedance, i.e. a constant scalar μ/ε ratio, with the possible exception of regions of perfectly conducting material.

More precisely, Feld-Tai reciprocity requires the Hermitian (or rather, complex-symmetric) symmetry of the electromagnetic operators as above, but also relies on the assumption that the operator relating and is a constant scalar multiple of the operator relating and which is true when ε is a constant scalar multiple of μ (the two operators generally differ by an interchange of ε and μ). As above, one can also construct a more general formulation for integrals over a finite volume.

Optical reciprocity in radiometric terms

Apart from quantal effects, classical theory covers near-, middle-, and far-field electric and magnetic phenomena with arbitrary time courses. Optics refers to far-field nearly-sinusoidal oscillatory electromagnetic effects. Instead of paired electric and magnetic variables, optics, including optical reciprocity, can be expressed in polarization-paired radiometric variables, such as spectral radiance, traditionally called specific intensity.

In 1856, Hermann von Helmholtz wrote:

"A ray of light proceeding from point A arrives at point B after suffering any number of refractions, reflections, &c. At point A let any two perpendicular planes a1, a2 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 like planes b1, b2 in the ray at point B; then the following proposition may be demonstrated. If when the quantity of light J polarized in the plane a1 proceeds from A in the direction of the given ray, that part K thereof of light polarized in b1 arrives at B, then, conversely, if the quantity of light J polarized in b1 proceeds from B, the same quantity of light K polarized in a1 will arrive at A." [13]

This is sometimes called the Helmholtz reciprocity (or reversion) principle. [14] [15] [16] [17] [18] [19] When the wave propagates through a material acted upon by an applied magnetic field, reciprocity can be broken so this principle will not apply. [13] Similarly, when there are moving objects in the path of the ray, the principle may be entirely inapplicable. Historically, in 1849, Sir George Stokes stated his optical reversion principle without attending to polarization. [20] [21] [22]

Like the principles of thermodynamics, 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. [23] [24]

The simplest statement of the principle is if I can see you, then you can see me. The principle was used by Gustav Kirchhoff in his derivation of his law of thermal radiation and by Max Planck in his analysis of his law of thermal radiation.

For ray-tracing global illumination algorithms, incoming and outgoing light can be considered as reversals of each other, without affecting the bidirectional reflectance distribution function (BRDF) outcome. [24]

Green's reciprocity

Whereas the above reciprocity theorems were for oscillating fields, Green's reciprocity is an analogous theorem for electrostatics with a fixed distribution of electric charge (Panofsky and Phillips, 1962).

In particular, let denote the electric potential resulting from a total charge density . The electric potential satisfies Poisson's equation, , where is the vacuum permittivity. Similarly, let denote the electric potential resulting from a total charge density , satisfying . In both cases, we assume that the charge distributions are localized, so that the potentials can be chosen to go to zero at infinity. Then, Green's reciprocity theorem states that, for integrals over all space:

This theorem is easily proven from Green's second identity. Equivalently, it is the statement that

i.e. that is a Hermitian operator (as follows by integrating by parts twice).

See also

Related Research Articles

<span class="mw-page-title-main">Maxwell's equations</span> Equations describing classical electromagnetism

Maxwell's equations, or Maxwell–Heaviside equations, are a set of coupled partial differential equations that, together with the Lorentz force law, form the foundation of classical electromagnetism, classical optics, electric and magnetic circuits. The equations provide a mathematical model for electric, optical, and radio technologies, such as power generation, electric motors, wireless communication, lenses, radar, etc. They describe how electric and magnetic fields are generated by charges, currents, and changes of the fields. The equations are named after the physicist and mathematician James Clerk Maxwell, who, in 1861 and 1862, published an early form of the equations that included the Lorentz force law. Maxwell first used the equations to propose that light is an electromagnetic phenomenon. The modern form of the equations in their most common formulation is credited to Oliver Heaviside.

In vector calculus and differential geometry the generalized Stokes theorem, also called the Stokes–Cartan theorem, is a statement about the integration of differential forms on manifolds, which both simplifies and generalizes several theorems from vector calculus. In particular, the fundamental theorem of calculus is the special case where the manifold is a line segment, Green’s theorem and Stokes' theorem are the cases of a surface in or and the divergence theorem is the case of a volume in Hence, the theorem is sometimes referred to as the Fundamental Theorem of Multivariate Calculus.

<span class="mw-page-title-main">Navier–Stokes equations</span> Equations describing the motion of viscous fluid substances

The Navier–Stokes equations are partial differential equations which describe the motion of viscous fluid substances. They were named after French engineer and physicist Claude-Louis Navier and the Irish physicist and mathematician George Gabriel Stokes. They were developed over several decades of progressively building the theories, from 1822 (Navier) to 1842–1850 (Stokes).

<span class="mw-page-title-main">Poynting vector</span> Measure of directional electromagnetic energy flux

In physics, the Poynting vector represents the directional energy flux or power flow of an electromagnetic field. The SI unit of the Poynting vector is the watt per square metre (W/m2); kg/s3 in base SI units. It is named after its discoverer John Henry Poynting who first derived it in 1884. Nikolay Umov is also credited with formulating the concept. Oliver Heaviside also discovered it independently in the more general form that recognises the freedom of adding the curl of an arbitrary vector field to the definition. The Poynting vector is used throughout electromagnetics in conjunction with Poynting's theorem, the continuity equation expressing conservation of electromagnetic energy, to calculate the power flow in electromagnetic fields.

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

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

<span class="mw-page-title-main">Gauss's law</span> Foundational law of electromagnetism relating electric field and charge distributions

In physics, Gauss's law, also known as Gauss's flux theorem, is one of Maxwell's equations. It relates the distribution of electric charge to the resulting electric field.

In vector calculus, the divergence theorem, also known as Gauss's theorem or Ostrogradsky's theorem, is a theorem relating the flux of a vector field through a closed surface to the divergence of the field in the volume enclosed.

<span class="mw-page-title-main">Holonomy</span> Concept in differential geometry

In differential geometry, the holonomy of a connection on a smooth manifold is a general geometrical consequence of the curvature of the connection measuring the extent to which parallel transport around closed loops fails to preserve the geometrical data being transported. For flat connections, the associated holonomy is a type of monodromy and is an inherently global notion. For curved connections, holonomy has nontrivial local and global features.

In mathematics, the total variation identifies several slightly different concepts, related to the (local or global) structure of the codomain of a function or a measure. For a real-valued continuous function f, defined on an interval [a, b] ⊂ R, its total variation on the interval of definition is a measure of the one-dimensional arclength of the curve with parametric equation xf(x), for x ∈ [a, b]. Functions whose total variation is finite are called functions of bounded variation.

In physics and mathematics, in the area of vector calculus, Helmholtz's theorem, also known as the fundamental theorem of vector calculus, states that any sufficiently smooth, rapidly decaying vector field in three dimensions can be resolved into the sum of an irrotational (curl-free) vector field and a solenoidal (divergence-free) vector field; this is known as the Helmholtz decomposition or Helmholtz representation. It is named after Hermann von Helmholtz.

In electrodynamics, Poynting's theorem is a statement of conservation of energy for electromagnetic fields developed by British physicist John Henry Poynting. It states that in a given volume, the stored energy changes at a rate given by the work done on the charges within the volume, minus the rate at which energy leaves the volume. It is only strictly true in media which is not dispersive, but can be extended for the dispersive case. The theorem is analogous to the work-energy theorem in classical mechanics, and mathematically similar to the continuity equation.

<span class="mw-page-title-main">Magnetic vector potential</span> Integral of the magnetic field

In classical electromagnetism, magnetic vector potential is the vector quantity defined so that its curl is equal to the magnetic field: . Together with the electric potential φ, the magnetic vector potential can be used to specify the electric field E as well. Therefore, many equations of electromagnetism can be written either in terms of the fields E and B, or equivalently in terms of the potentials φ and A. In more advanced theories such as quantum mechanics, most equations use potentials rather than fields.

The electric-field integral equation is a relationship that allows the calculation of an electric field generated by an electric current distribution.

The electromagnetic wave equation is a second-order partial differential equation that describes the propagation of electromagnetic waves through a medium or in a vacuum. It is a three-dimensional form of the wave equation. The homogeneous form of the equation, written in terms of either the electric field E or the magnetic field B, takes the form:

In differential calculus, the Reynolds transport theorem, or simply the Reynolds theorem, named after Osborne Reynolds (1842–1912), is a three-dimensional generalization of the Leibniz integral rule. It is used to recast time derivatives of integrated quantities and is useful in formulating the basic equations of continuum mechanics.

In statistics, generalized least squares (GLS) is a method used to estimate the unknown parameters in a linear regression model. It is used when there is a non-zero amount of correlation between the residuals in the regression model. GLS is employed to improve statistical efficiency and reduce the risk of drawing erroneous inferences, as compared to conventional least squares and weighted least squares methods. It was first described by Alexander Aitken in 1935.

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

Multipole radiation is a theoretical framework for the description of electromagnetic or gravitational radiation from time-dependent distributions of distant sources. These tools are applied to physical phenomena which occur at a variety of length scales - from gravitational waves due to galaxy collisions to gamma radiation resulting from nuclear decay. Multipole radiation is analyzed using similar multipole expansion techniques that describe fields from static sources, however there are important differences in the details of the analysis because multipole radiation fields behave quite differently from static fields. This article is primarily concerned with electromagnetic multipole radiation, although the treatment of gravitational waves is similar.

<span class="mw-page-title-main">Loop representation in gauge theories and quantum gravity</span> Description of gauge theories using loop operators

Attempts have been made to describe gauge theories in terms of extended objects such as Wilson loops and holonomies. The loop representation is a quantum hamiltonian representation of gauge theories in terms of loops. The aim of the loop representation in the context of Yang–Mills theories is to avoid the redundancy introduced by Gauss gauge symmetries allowing to work directly in the space of physical states. The idea is well known in the context of lattice Yang–Mills theory. Attempts to explore the continuous loop representation was made by Gambini and Trias for canonical Yang–Mills theory, however there were difficulties as they represented singular objects. As we shall see the loop formalism goes far beyond a simple gauge invariant description, in fact it is the natural geometrical framework to treat gauge theories and quantum gravity in terms of their fundamental physical excitations.

In optics, the Ewald–Oseen extinction theorem, sometimes referred to as just the extinction theorem, is a theorem that underlies the common understanding of scattering. It is named after Paul Peter Ewald and Carl Wilhelm Oseen, who proved the theorem in crystalline and isotropic media, respectively, in 1916 and 1915. Originally, the theorem applied to scattering by an isotropic dielectric objects in free space. The scope of the theorem was greatly extended to encompass a wide variety of bianisotropic media.

References

Citations

  1. Stumpf, M. (2018). Electromagnetic Reciprocity in Antenna Theory. Piscataway, NJ: Wiley-IEEE Press.
  2. Balanis, C. A. (2024). Advanced Engineering Electromagnetics (3 ed.). Wiley. pp. 335–337. doi:10.1002/9781394180042. ISBN   978-1-394-18001-1 . Retrieved February 29, 2024.
  3. Harrington, R. F. (2001). Time-Harmonic Electromagnetic Fields. Wiley-IEEE. pp. 116–120. doi:10.1109/9780470546710. ISBN   978-0-471-20806-8 . Retrieved February 29, 2024.
  4. Van Bladel, J. (2007). Electromagnetic Fields (2 ed.). Wiley-IEEE. pp. 397–402. doi:10.1002/047012458X. ISBN   978-0-471-26388-3 . Retrieved February 29, 2024.
  5. Felsen, L. B.; Marcuvitz, N. (2003). Radiation and Scattering of Waves. Wiley-IEEE. pp. 90–93. doi:10.1109/9780470546307. ISBN   978-0-780-31088-9 . Retrieved February 29, 2024.
  6. Collin, R. E. (1991). Field Theory of Guided Waves (2 ed.). Wiley-IEEE. pp. 49–50. doi:10.1109/9780470544648. ISBN   978-0-879-42237-0 . Retrieved February 29, 2024.
  7. Chew, Wen Cho (April 2008). "A new look at reciprocity and energy conservation theorems in electromagnetics". IEEE Transactions on Antennas and Propagation. 56 (4): 970–975. Bibcode:2008ITAP...56..970C. doi:10.1109/TAP.2008.919189. S2CID   13615400.
  8. Ramo, Simon; Whinnery, John; van Duzer, Theodore (1965). Fields and Waves in Communication Electronics (International ed.). John Wiley & Sons. ISBN   978-047170720-2. ISBN   0471707201
  9. Stumpf, M. (2018). Electromagnetic Reciprocity in Antenna Theory. IEEE Press / Wiley. §1.4.3.
  10. Kong, Jin Au (1972). "Theorems of bianisotropic media". Proceedings of the IEEE. 60 (9): 1036–1046. doi:10.1109/PROC.1972.8851.
  11. Feld, Ya.N. (1992). "On the quadratic lemma in electrodynamics". Sov. Phys. Dokl. 37: 235–236.
  12. Tai, C.-T. (1992). "Complementary reciprocity theorems in electromagnetic theory". IEEE Transactions on Antennas and Propagation. 40 (6): 675–681. Bibcode:1992ITAP...40..675T. doi:10.1109/8.144602. hdl: 2027.42/21036 .
  13. 1 2 von Helmholtz, H. (1856). Handbuch der physiologischen Optik[Handbook of Physiological Optics]. Vol. 1 (1st ed.). Leipzig: Leopold Voss. p. 169; cited by Planck. English version quoted here based on Translated by Guthrie, F. "translation of Helmholtz". Philosophical Magazine (Second printing). Series 4. 20: 2–21. 1867.
  14. Minnaert, M. (1941). "The reciprocity principle in lunar photometry". The Astrophysical Journal . 93: 403–410. Bibcode:1941ApJ....93..403M. doi:10.1086/144279.
  15. Chandrasekhar, S. (1950). Radiative Transfer. Oxford, UK: Oxford University Press. pp. 20–21, 171–177, 182.
  16. Tingwaldt, C.P. (1952). "Über das Helmholtzsche Reziprozitätsgesetz in der Optik" [On the Helmholtz law of reciprocity in optics]. Optik. 9 (6): 248–253.
  17. Levi, L. (1968). Applied Optics: A guide to optical system design. Vol. 1. New York, NY: Wiley. p. 84. (2 vols.)
  18. Clarke, F.J.J.; Parry, D.J. (1985). "Helmholtz reciprocity: Its validity and application to reflectometry". Lighting Research & Technology . 17 (1): 1–11. doi:10.1177/14771535850170010301. S2CID   123394330.
  19. Born, M.; Wolf, E. (1999). Principles of Optics: Electromagnetic theory of propagation, interference and diffraction of light (7th ed.). Cambridge University Press. p. 423. ISBN   0-521-64222-1.
  20. 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.
  21. Mahan, A.I. (1943). "A mathematical proof of Stokes' reversibility principle". Journal of the Optical Society of America . 33 (11): 621–626. doi:10.1364/JOSA.33.000621.
  22. Lekner, J. (1987). Theory of Reflection of Electromagnetic and Particle Waves. Dordrecht: Martinus Nijhoff. pp. 33–37. ISBN   90-247-3418-5 via Google Books.
  23. Rayleigh, J.W. Strutt, baron (1900). "On the law of reciprocity in diffuse reflection". Philosophical Magazine . series 5. 49: 324–325.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  24. 1 2 Hapke, B. (1993). Theory of Reflectance and Emittance Spectroscopy. Cambridge UK: Cambridge University Press. Section 10C, pages 263-264. ISBN   0-521-30789-9.