Fact-checked by Grok 2 weeks ago

Kirchhoff integral theorem

The Kirchhoff integral theorem, also known as the Kirchhoff-Helmholtz integral theorem, is a foundational result in classical wave theory that provides an integral representation for solutions to the scalar (\nabla^2 + k^2) u = 0 within a volume V enclosed by a surface S, expressing the wave function u(\mathbf{r}) at any interior point \mathbf{r} in terms of the boundary values of u and its \partial u / \partial n on S. The theorem is derived from Green's second identity applied to u and the G(\mathbf{r}, \mathbf{r}') = \frac{e^{ik|\mathbf{r} - \mathbf{r}'|}}{4\pi |\mathbf{r} - \mathbf{r}'|}, yielding the formula
u(\mathbf{r}) = \oint_S \left[ G(\mathbf{r}, \mathbf{r}') \frac{\partial u}{\partial n'}(\mathbf{r}') - u(\mathbf{r}') \frac{\partial G}{\partial n'}(\mathbf{r}, \mathbf{r}') \right] dS',
where k = 2\pi / \lambda is the and the is outward-pointing. This representation assumes a source-free region inside V and monochromatic waves in a linear, isotropic medium.
Formulated by German physicist Gustav Robert Kirchhoff in his 1882 paper "Zur Theorie der Lichtstrahlen" presented to the , the theorem built upon earlier work by Fresnel on Huygens' principle and Stokes' application of to wave equations, providing a rigorous mathematical framework for phenomena. Kirchhoff's derivation involved integrating the wave equation over a volume bounded by an and opaque screen, imposing conditions where the wave vanishes on the screen and equals the incident in the aperture (known as Kirchhoff's -value ). Although the theory contains mathematical inconsistencies—such as the Poincaré , where the assumed conditions imply no wave propagation— it yields accurate predictions in the geometric limit and for small apertures, explaining its enduring influence despite later refinements by and Sommerfeld. The theorem has broad applications across wave physics, including for calculating diffracted light fields through via the Fresnel-Kirchhoff diffraction formula, which incorporates an obliquity (inclination) factor \frac{1 + \cos \chi}{2} to account for directional contributions from secondary wavelets:
u(\mathbf{r}) = \frac{1}{i\lambda} \iint_A u(\mathbf{r}') \frac{e^{iks}}{s} \frac{1 + \cos \chi}{2} dA',
where s is the distance from aperture point to , \chi the angle between incident and diffracted directions, and A the aperture. In acoustics, it underpins near-field acoustical and radiation calculations by relating and on a surface to the field everywhere. Extensions to yield vector forms for solving in scattering problems, while in , one-way approximations facilitate wavefield migration for imaging subsurface structures.

Mathematical Foundations

Scalar Wave Equation

The scalar wave equation is a second-order linear that governs the propagation of waves in a homogeneous, isotropic medium for s. It takes the form \nabla^2 u - \frac{1}{c^2} \frac{\partial^2 u}{\partial t^2} = 0, where u(\mathbf{r}, t) denotes the (such as displacement or potential), \mathbf{r} is the spatial position, t is time, \nabla^2 is the Laplacian, and c is the constant wave speed in the medium. This equation arises in contexts where wave phenomena can be approximated by a scalar quantity, neglecting vectorial complexities. Physically, the scalar wave equation models acoustic pressure variations in fluids, scalar potentials in electromagnetic theory under certain approximations, and optical fields via the scalar diffraction approximation in wave . For instance, in acoustics, u represents acoustic pressure, while in optics, it approximates the component perpendicular to the direction for paraxial beams. The equation traces its origins to Jean le Rond d'Alembert's 1747 derivation for the one-dimensional vibrating string, marking the birth of partial differential equations for wave motion. In the 19th century, it became central to wave optics, enabling rigorous treatments of and , as in Kirchhoff's 1882 integral theorem for boundary value problems. Assuming a time-harmonic dependence u(\mathbf{r}, t) = U(\mathbf{r}) e^{-i \omega t} (with the real part implied), where \omega is the , the equation simplifies to the time-independent Helmholtz form: \nabla^2 U + k^2 U = 0, with wavenumber k = \omega / c. This form facilitates frequency-domain analysis of steady-state wave propagation. Green's identities serve as key mathematical tools for deriving integral solutions to boundary value problems of this equation.

Green's Identities

Green's first identity relates the volume integral of a scalar function and its Laplacian to a surface integral involving its normal derivative. For two sufficiently smooth scalar functions u and v defined on a volume V bounded by a closed surface S, it states: \int_V \left( u \nabla^2 v + \nabla u \cdot \nabla v \right) dV = \oint_S u \frac{\partial v}{\partial n} \, dS, where \partial / \partial n = \mathbf{n} \cdot \nabla denotes the outward normal derivative on S. Green's second identity follows by subtracting the first identity with the roles of u and v interchanged, yielding: \int_V \left( u \nabla^2 v - v \nabla^2 u \right) dV = \oint_S \left( u \frac{\partial v}{\partial n} - v \frac{\partial u}{\partial n} \right) dS. This identity is fundamental for deriving integral representations of solutions to partial differential equations, such as the , by choosing an appropriate v that satisfies the homogeneous equation except at singularities. To derive these identities, apply the to the u \nabla v: \int_V \nabla \cdot (u \nabla v) \, dV = \oint_S u \frac{\partial v}{\partial n} \, dS. The left side expands using the \nabla \cdot (u \nabla v) = \nabla u \cdot \nabla v + u \nabla^2 v, directly giving the first identity. The second follows by subtraction, as the \nabla u \cdot \nabla v terms cancel. In the context of the , the auxiliary function v is chosen as the fundamental solution, which satisfies the equation away from the source point. For the time-dependent case, the retarded Green's function takes the form \frac{1}{4\pi s} \delta\left(t - t' - \frac{s}{c}\right), where s = |\mathbf{r} - \mathbf{r}'| is the and c is the wave speed; this ensures by incorporating the . For the monochromatic case, corresponding to the , v = \frac{e^{i k s}}{4\pi s}, where k is the , providing an outgoing spherical wave solution except at the .

Statement of the Theorem

Monochromatic Case

In the monochromatic case, the Kirchhoff integral theorem addresses time-harmonic wave fields that satisfy the homogeneous \nabla^2 U + k^2 U = 0 within a volume V bounded by a closed surface S, where k = 2\pi / \lambda is the and no sources are present inside V. This formulation assumes the field U and its normal are known on S, with the functions being sufficiently (continuous first- and second-order partial derivatives) on and within S. Far-field approximations may be invoked when distances are large compared to the , but the theorem holds exactly under the stated conditions without such approximations. For a point \mathbf{r} inside V, the field is given by the surface integral U(\mathbf{r}) = \frac{1}{4\pi} \int_S \left[ U \frac{\partial}{\partial \hat{\mathbf{n}}} \left( \frac{e^{i k s}}{s} \right) - \frac{e^{i k s}}{s} \frac{\partial U}{\partial \hat{\mathbf{n}}} \right] dS, where s = |\mathbf{r} - \mathbf{r}'| is the distance from the observation point \mathbf{r} to the integration point \mathbf{r}' on S, and \hat{\mathbf{n}} denotes the outward-pointing unit normal to S. This representation, derived from Green's second identity applied to the , expresses the interior field solely in terms of boundary values. The integral combines Dirichlet boundary conditions (values of U on S) with Neumann boundary conditions (values of \partial U / \partial \hat{\mathbf{n}} on S) to uniquely recover the field inside V, embodying Huygens' principle in the where secondary wavelets from the surface constructively interfere to reproduce the total field. As an implication, if both U = 0 and \partial U / \partial \hat{\mathbf{n}} = 0 on S, then U(\mathbf{r}) = 0 for all \mathbf{r} inside V, underscoring the uniqueness of solutions to the homogeneous under specified boundary conditions.

Time-Dependent Case

The time-dependent case of the Kirchhoff integral theorem extends the representation to solutions of the scalar that vary arbitrarily with time, applicable to broadband or pulsed wave phenomena such as transient acoustic or electromagnetic disturbances. This formulation arises from applying Green's second identity to the \square V = 0, where \square = \partial^2 / \partial t^2 - c^2 \nabla^2 is the d'Alembertian operator and c is the constant propagation speed (e.g., or ). The theorem provides the value of the field V at an interior point \mathbf{r} and time t solely in terms of boundary data on a closed surface S enclosing \mathbf{r}, assuming no sources inside the volume bounded by S. Under these assumptions—homogeneity of the wave equation within the volume, a closed orientable surface S excluding sources, and constant speed c—the theorem states that the solution is given by the surface integral V(\mathbf{r}, t) = \frac{1}{4\pi} \int_S \left\{ \frac{1}{s} \left[ \frac{\partial V}{\partial n} \right] - [V] \frac{\partial}{\partial n} \left( \frac{1}{s} \right) + \frac{1}{c s} \frac{\partial s}{\partial n} \left[ \frac{\partial V}{\partial t} \right] \right\} dS, where s = |\mathbf{r} - \mathbf{r}'| is the distance from the observation point \mathbf{r} to the surface point \mathbf{r}', \partial / \partial n denotes the outward normal derivative on S, and the square brackets denote retarded evaluation: = f(t - s/c, \mathbf{r}'). This expression incorporates contributions from the field V, its normal derivative, and its time derivative on S, weighted by geometric factors involving $1/s and its derivatives. The retarded arguments ensure causality, as the field at (\mathbf{r}, t) depends only on boundary values emitted at earlier times t - s/c, reflecting the finite propagation speed c and preventing instantaneous influences across distances. This time-delay mechanism distinguishes the time-dependent theorem from steady-state formulations, capturing dispersive effects in transient waves. In the static limit, where c \to \infty (or equivalently, for time-independent fields), the retardation vanishes (s/c \to 0), the third term drops out, and the formula reduces to Poisson's integral representation for solutions of \nabla^2 V = 0: V(\mathbf{r}) = \frac{1}{4\pi} \int_S \left( \frac{1}{s} \frac{\partial V}{\partial n} - V \frac{\partial}{\partial n} \left( \frac{1}{s} \right) \right) dS. Thus, the time-dependent case generalizes the electrostatic or steady-state to dynamic scenarios while preserving consistency in the quasistatic regime.

Derivation

Application of Green's Second Identity

To derive the Kirchhoff integral theorem, Green's second identity is applied to a U satisfying the homogeneous within a volume V bounded by surface S, with the observation point \mathbf{r} located inside V. In the monochromatic case, the is chosen as \psi(\mathbf{r}', \mathbf{r}) = \frac{e^{ik s}}{s}, where s = |\mathbf{r}' - \mathbf{r}| and k is the ; this function satisfies the (\nabla^2 + k^2) \psi = 0 everywhere except at the \mathbf{r}' = \mathbf{r}. For the time-dependent case, the is the \psi(\mathbf{r}', \mathbf{r}, t, t') = \frac{\delta(t - t' - s/c)}{s}, which satisfies the \nabla^2 \psi - \frac{1}{c^2} \frac{\partial^2 \psi}{\partial t^2} = 0 away from the . Green's second identity states that for sufficiently smooth functions U and \psi, \int_V \left( U \nabla^2 \psi - \psi \nabla^2 U \right) dV = \oint_S \left( U \frac{\partial \psi}{\partial n} - \psi \frac{\partial U}{\partial n} \right) dS, where \partial / \partial n denotes the outward normal derivative on S. Away from the at \mathbf{r}, both U and \psi satisfy their respective homogeneous equations—in the monochromatic case, \nabla^2 U + k^2 U = 0 and \nabla^2 \psi + k^2 \psi = 0, so the integrand simplifies to U (\nabla^2 \psi) - \psi (\nabla^2 U) = U (-k^2 \psi) - \psi (-k^2 U) = 0; similarly, in the time-dependent case, the terms vanish away from \mathbf{r}. Thus, the volume integral vanishes everywhere except in a small neighborhood of the at \mathbf{r}. This reduction implies that U(\mathbf{r}) can be expressed solely in terms of the values of U and its normal derivative on the bounding surface S, providing the foundation for the surface representation central to the theorem.

Handling the

In the derivation of the Kirchhoff integral theorem, the auxiliary Green's function is \psi(\mathbf{r}, \mathbf{r}') = \frac{e^{ik|\mathbf{r} - \mathbf{r}'|}}{|\mathbf{r} - \mathbf{r}'|}\ ) for the monochromatic case (without the \(1/4\pi here for clarity in singularity ), introducing a singularity at \mathbf{r} = \mathbf{r}'. To resolve this, the observation point is excluded from the integration volume by surrounding it with a small of \epsilon, ensuring the functions remain smooth within the modified . Green's second identity is then applied to the region between the primary surface S and this small sphere, transforming the volume integral into surface contributions over both boundaries. The normal on the small sphere points outward from the volume (inward toward \mathbf{r}), which introduces a sign flip relative to the radial derivative. The contribution from the small sphere is evaluated using the divergence theorem on the integrand U \nabla \psi - \psi \nabla U, where U is the wave field satisfying the Helmholtz equation. As \epsilon \to 0, U and its gradient are approximately constant over the sphere, while \psi \approx 1/\epsilon and \partial \psi / \partial n \approx 1/\epsilon^2 (accounting for the inward normal sign). The surface integral over the small sphere thus simplifies to \int_{\text{small}} (U \partial \psi / \partial n - \psi \partial U / \partial n) \, dS \approx -4\pi U(\mathbf{r}), with the second term vanishing due to its O(\epsilon) scaling. This isolates the desired U(\mathbf{r}) term, confirming the factor of $4\pi in the final representation. In the limiting process, the surface integral over the primary S (with outward ) equates to $4\pi U(\mathbf{r}) after adding the small sphere's contribution (with appropriate ), yielding U(\mathbf{r}) = \frac{1}{4\pi} \oint_S \left[ U \frac{\partial \psi}{\partial n'} - \psi \frac{\partial U}{\partial n'} \right] dS'. For the time-dependent case, the auxiliary function is the \psi(\mathbf{r}, \mathbf{r}', t - t') = \frac{\delta(t - t' - s/c)}{s}, and the small sphere analysis proceeds analogously, with retarded times on approaching the observation time uniformly as \epsilon \to 0, preserving the $4\pi factor. This approach ensures the theorem's validity across both formulations.

Applications

Diffraction Theory

In diffraction theory, the Kirchhoff integral theorem provides a foundational framework for predicting the propagation of scalar waves through apertures in opaque screens, approximating the behavior of light under the scalar . Kirchhoff's seminal 1882 work, published in 1883, extended the empirical Huygens-Fresnel by deriving a rigorous solution based on , thereby establishing a mathematical basis for classical phenomena. Central to this application are Kirchhoff's boundary conditions, which assume that the incident wave field passes unchanged through the while vanishing, along with its normal derivative, on the opaque portions of the screen; these conditions are approximate and idealize the transition at the aperture edges. Applying the monochromatic case of the to an observation point P beyond the screen yields the Fresnel-Kirchhoff diffraction formula: U(P) = \frac{1}{4\pi} \int_{\text{aperture}} \left( U \frac{\partial}{\partial n} \left( \frac{e^{ikr}}{r} \right) - \frac{\partial U}{\partial n} \frac{e^{ikr}}{r} \right) dS, where U is the field over the aperture (taken as the incident field), r is the distance from aperture elements to P, k = 2\pi/\lambda is the wavenumber, and n denotes the normal derivative; an obliquity factor, often (1 + \cos \theta)/2 with \theta the angle between the normal and line to P, is incorporated to account for directional propagation. This formula underpins both near-field (Fresnel) and far-field ( regimes. In the Fresnel , valid when the observation distance z satisfies z \ll a^2 / \lambda (with a the size), the integral captures curved wavefronts and zone-plate effects. For , at large z where z \gg a^2 / \lambda, the formula simplifies to a of the function, yielding linear intensity patterns independent of z. A representative example is single-slit , where the Fraunhofer pattern exhibits a central maximum with minima at angles \sin \theta = m \lambda / a (m = \pm 1, \pm 2, \dots), demonstrating from secondary wavelets across the slit width.

Acoustic Propagation

In acoustics, the Kirchhoff integral theorem applies to the propagation of waves, where the acoustic variable U represents the pressure deviation from ambient conditions, satisfying the homogeneous scalar \nabla^2 U - \frac{1}{c^2} \frac{\partial^2 U}{\partial t^2} = 0 in source-free , with c denoting the in the medium. The theorem expresses the interior acoustic field within an enclosed or as surface and integrals over data, such as and measured by arrays on enclosure walls or surrounding surfaces, enabling reconstruction of the field from partial observations. This formulation is particularly useful for analyzing in bounded spaces like rooms or automotive interiors, where conditions on walls dictate the interior distribution. A key application involves sound radiation from vibrating surfaces, such as panels in machinery or bodies excited by transient forces; here, the time-dependent form of the theorem incorporates retarded potentials to capture delays, allowing prediction of the radiated acoustic in open or semi-enclosed environments. For transient sources, this approach facilitates the simulation of impulsive noises, like those from impacts, by integrating over the vibrating surface's and distributions. Numerically, the (BEM) implements the theorem by discretizing the surface into panels, converting the continuous integral into a matrix equation solved for unknown boundary values, which is efficient for low-frequency acoustic simulations in complex geometries like outdoor propagation or reverberant spaces. BEM reduces the problem dimensionality compared to volume-based methods, making it suitable for engineering predictions of noise radiation. In free-field conditions, the theorem with the free-field —representing outgoing spherical waves from point sources—enables recovery of the acoustic field generated by monopolar sources, such as engine exhausts, supporting applications in near-field acoustical for source localization and strength estimation.

Limitations and Extensions

Key Assumptions and Inconsistencies

The Kirchhoff integral theorem relies on several foundational assumptions to derive the solution to the homogeneous scalar within a volume bounded by a closed surface. Central to these is the requirement of a closed surface enclosing the observation point, with no sources present inside this volume, ensuring that the wave field satisfies the without external excitations. Additionally, the theorem assumes the validity of specific conditions, such as the wave function and its normal derivative being zero on opaque portions of the , which implies discontinuous fields at edges or apertures. These assumptions facilitate the application of Green's second identity but introduce idealizations that do not fully align with physical reality. A primary mathematical inconsistency arises from Kirchhoff's boundary conditions, which overdetermine the problem by simultaneously specifying both the wave function and its normal derivative on the boundary, violating the uniqueness theorem for solutions to the wave equation. This leads to a paradox, as noted by Poincaré in 1892, where the derived field does not recover the assumed boundary values when evaluated back on the surface itself. Furthermore, the theorem ignores contributions from edge diffraction, treating the aperture as a continuous source of secondary wavelets without accounting for disturbances at the boundaries, which results in significant errors in the near-field region close to apertures or obstacles. Physically, the scalar formulation of the theorem neglects vectorial effects such as , limiting its applicability to unpolarized or scalar waves and failing to capture the full behavior of electromagnetic fields. It also presupposes a far-field where the s from the to the observation point greatly exceeds the \lambda (i.e., s \gg \lambda), along with the aperture dimensions being large compared to \lambda, which restricts accuracy in near-field or low-frequency scenarios. These limitations were critiqued historically by Lord Rayleigh in 1897, who highlighted the unphysical nature of Kirchhoff's boundary conditions and proposed alternative formulations that specify only one condition per region to restore consistency. Similarly, in 1896 developed a rigorous approach using Green's functions for semi-infinite screens, exposing the theorem's failure to satisfy the wave equation at boundaries and motivating exact methods that incorporate edge effects more accurately. Despite these flaws, the theorem's practical success in predicting patterns stems from the wave equation's tolerance for such approximations in many optical contexts.

Generalizations to Other Fields

The Kirchhoff integral theorem, originally formulated for scalar wave fields, has been extended to vector electromagnetic fields through the Stratton-Chu formulas, which express the \mathbf{E} and \mathbf{H} as surface integrals over a closed surface enclosing the observation point. These formulas incorporate tangential components of the fields on the surface and employ Green's functions to account for the nature of the waves, deriving from vector generalizations of Green's second identity. The approach resolves inconsistencies in earlier vector diffraction theories by ensuring equivalence between formulations based on different vector identities. In , analogous integral representations appear in theory, where semiclassical approximations yield formulations that mirror the Kirchhoff theorem's superposition of secondary waves. These representations describe wave propagation and in quantum systems, such as from apertures, by integrating over paths akin to the Huygens-Kirchhoff principle in . Such analogies facilitate semiclassical treatments of quantum , linking classical wave to quantum path sums in problems. Computational extensions leverage fast multipole methods (FMM) to accelerate evaluation of the Kirchhoff-Helmholtz integrals for large-scale problems, reducing complexity from O(N^2) to O(N) or O(N \log N) by hierarchical expansions of the . These methods are particularly effective in boundary element formulations for acoustic scattering, enabling simulations of complex geometries with millions of elements. Time-domain implementations of the Kirchhoff integral, derived via transforms of the frequency-domain version, support transient acoustic predictions by integrating retarded potentials over moving surfaces, avoiding frequency-by-frequency computations. This formulation simplifies handling of convective effects in uniform flows, with improved through specialized rules. Post-2000 advancements in computational acoustics have integrated the Kirchhoff integral into advanced boundary element methods (BEM) for high-frequency and , incorporating adaptive fast multipole acceleration for ultra-large models up to $10^7 elements. Relativistic generalizations extend the theorem to curved spacetimes, applying Kirchhoff integrals to scalar fields in geometries to compute Green's functions and wave propagation, ensuring covariance under Lorentz transformations. These developments enhance simulations of aeroacoustic noise from rotating machinery and flow-generated sound, combining time-domain Kirchhoff formulations with hybrid finite-volume/BEM approaches for broadband predictions.

References

  1. [1]
    [PDF] Chapter 15S Fresnel-Kirchhoff diffraction - bingweb
    Jan 22, 2011 · Fresnel-Kirchhoff diffraction theory​​ According to the Huygen's construction, every point of a wave-front may be considered as a center of a ...
  2. [2]
    None
    Summary of each segment:
  3. [3]
    The Kirchhoff–Helmholtz integral theorem and related identities for ...
    Apr 1, 1996 · The KHT is applied to study uniqueness of solutions to acoustic boundary value problems in moving media and to establish unitarity and other ...
  4. [4]
    1.5: The Scalar and Vector Wave Equation - Physics LibreTexts
    Sep 16, 2022 · In a homogeneous dielectric without external sources, every component of the electromagnetic field satisfies the scalar wave equation.
  5. [5]
    Wave Equation - an overview | ScienceDirect Topics
    The wave equation is defined as a partial differential equation that describes the behavior of various types of waves, including sound, light, and water ...
  6. [6]
    D'Alembert and the Wave Equation: Its Disputes and Controversies
    The concept of partial differential equation was introduced by d'Alembert in about 1740 to solve problems of continuous media. He recognized in this field a new ...
  7. [7]
    1.6: Time-Harmonic Solutions of the Wave Equation
    Sep 16, 2022 · We define T = 2π/ω and λ = 2π/k as the period and the wavelength in the material, respectively. Furthermore, λ0 = 2π/k0 is the wavelength in ...
  8. [8]
    [PDF] Springer Undergraduate Mathematics Series
    Vector calculus. - (Springer undergraduate mathematics series). 1. Vector ... This result is known as Green's First Identity. • Choose u = IV g- gV f ...<|control11|><|separator|>
  9. [9]
    [PDF] Green's Identities
    Equation (21.6) is known as Green's first identity. ( r , r ) ndS : (21.8) Equation (21.8) is known as Green's second identity. Equation (21.9) is known as ...
  10. [10]
    Retarded potentials - Richard Fitzpatrick
    Every charge and every current in the Universe emits these spherical waves. The resultant scalar and vector potential fields are given by Eqs. (512) and (513).
  11. [11]
  12. [12]
    [PDF] 19690014676.pdf - NASA Technical Reports Server (NTRS)
    The derivation of the scalar Kirchhoff integral theorem involves several assumptions and the limitations of the theory, which become evident In the derivation, ...
  13. [13]
    [PDF] 7.7 Kirchhoff Theory
    Kirchhoff theory can be used to provide an approximate solution for either the transmitted or reflected wavefield due to this interface (Fig. 7.18).Missing: explanation | Show results with:explanation
  14. [14]
    A time-domain Kirchhoff formula for the convective acoustic wave ...
    Mar 1, 2016 · Kirchhoff's integral method allows propagated sound to be predicted, based on the pressure and its derivatives in time and space obtained on ...
  15. [15]
    [PDF] Time-Reversed Diffraction 1 Problem 2 The Kirchhoff Integral via ...
    In the usual formulation of the Kirchhoff diffraction integral [1], a scalar field with harmonic time dependence at angular frequency ω is deduced in a charge/ ...
  16. [16]
    [PDF] Acoustic Wave - FSU Mathematics
    According to Green's Second Identity: (. ) (. ) ∫ ∫∫. ∫∫∫∫. +. + ... The Kirchhoff's formula, also called the Kirchhoff-Helmholtz Integral Theorem, was first.
  17. [17]
    [PDF] Transfer Characteristics of White Light Interferometers and Confocal ...
    Jul 10, 2017 · At first, a brief derivation of the Kirchhoff integral theorem is given. We assume U is the space-dependent part of a monochromatic scalar ...<|control11|><|separator|>
  18. [18]
    Zur Theorie der Lichtstrahlen - Kirchhoff - 1883 - Wiley Online Library
    1882. Google Scholar. p693_1) Es wird ohne Schwierigkeit sich ... Download PDF. back. Additional links. About Wiley Online Library. Privacy Policy ...
  19. [19]
    [PDF] Introduction to Modern Optics - CLASSE (Cornell)
    This equation is known as the Kirchhoff integral theorem. It relates the value of any scalar wave function at any point P inside an arbitrary closed surface ...
  20. [20]
    [PDF] Vectorizing Green's identities
    May 5, 2021 · ... Green's second vector identity and via the Kirchhoff formula. The derivations are unnecessarily restricted to a Cartesian coordinate system ...
  21. [21]
    Stratton–Chu vectorial diffraction of electromagnetic fields by ...
    The Stratton–Chu theory of electromagnetic (EM) scattering is used to develop a Kirchhoff formalism of the diffraction of EM waves by an aperture.
  22. [22]
    [PDF] Semiclassical approximations in wave mechanics
    has shown how to form similar 'semiclassical integral representations' for very general problems; they are analogous to Kirchhoff's integral in physical optics.
  23. [23]
    Quantum-mechanical diffraction theory of light from a small hole
    Jul 27, 2015 · In Sec. IV A , we start from a 2D electron version of the famous optical (acoustic) integral theorem of Helmholtz and Kirchhoff
  24. [24]
    Novel and efficient implementation of multi-level fast multipole ...
    This paper presents a novel formulation of the multi-level fast multipole indirect BEM, which relies on redesigning a conventional FMBEM with considerations ...<|separator|>
  25. [25]
    Time-domain Helmholtz-Kirchhoff integral for surface scattering in a ...
    Mar 14, 2017 · The Helmholtz-Kirchhoff (H-K) integral is widely used to study scattering of acoustic waves from rough surfaces,1,2 especially when the acoustic ...Introduction · Time-domain H-K integral in a... · H-K integral · Numerical resultsMissing: theorem | Show results with:theorem
  26. [26]
    [PDF] Solution of Kirchhoff's Time-Domain Integral Equation in Acoustics
    Mar 29, 2005 · In this paper we will look at the time-domain equivalent of the Helmholtz integral equation in the frequency domain. This integral equation ...
  27. [27]
    [PDF] arXiv:1204.0407v2 [gr-qc] 11 Apr 2012
    Apr 11, 2012 · In this paper, we apply the Kirchhoff integral method to the case of a scalar field propagating in a black hole toy-model spacetime M = M2 × S2, ...
  28. [28]
    (PDF) Recent Advances in Acoustic Boundary Element Methods
    Oct 10, 2022 · This paper reviews what is commonly known as direct BEM for linear time-harmonic acoustics. After introducing the boundary integral formulation ...
  29. [29]
    Computational Prediction of Flow-Generated Sound - Annual Reviews
    Abstract. This article provides a critical review of computational techniques for flow-noise prediction and the underlying theories.