Fact-checked by Grok 2 weeks ago

Kirchhoff's diffraction formula

Kirchhoff's diffraction formula is a fundamental equation in scalar diffraction theory that approximates the propagation of light waves through an in an opaque screen, expressing at an observation point as a over the aperture involving the incident , its normal , and a for spherical waves. Developed by the German Gustav Robert Kirchhoff in , it builds on earlier ideas from Huygens' principle and Fresnel's zone construction to provide a rigorous mathematical derivation using applied to the Helmholtz . The formula assumes a monochromatic scalar wave , with the amplitude and its normal vanishing on the screen outside the aperture—known as Kirchhoff's boundary conditions—and incorporates an obliquity factor to account for the directional dependence of secondary wavelets. Historically, Kirchhoff's work addressed limitations in Fresnel's empirical approach by integrating it with the wave equation, though it predates full electromagnetic theory and relies on scalar approximations valid for wavelengths much smaller than aperture dimensions. Despite its mathematical inconsistencies—such as the over-specification of conditions leading to Poincaré's in 1889—the formula yields accurate predictions for far-field (Fraunhofer) and near-field ( patterns, making it a of optical analysis. Subsequent refinements, including the Rayleigh-Sommerfeld diffraction integrals in 1896 and 1909, relaxed these conditions for greater physical consistency while preserving the core integral structure. In practice, the Fresnel-Kirchhoff form simplifies the general for paraxial approximations, enabling computations of distributions in applications like , , and antenna design, where it effectively models from edges. The key expression is u(\mathbf{r}) = \frac{1}{4\pi} \iint_A \left[ u(\mathbf{r}') \frac{\partial G}{\partial n'} - G \frac{\partial u}{\partial n'} \right] dA', with G as the outgoing \frac{e^{ik|\mathbf{r}-\mathbf{r}'|}}{|\mathbf{r}-\mathbf{r}'|}, highlighting the balance between field values and their gradients across the . Though superseded by vectorial theories for polarized , its enduring utility stems from empirical success and foundational role in education.

Historical and conceptual background

Development and context

The development of Kirchhoff's diffraction formula emerged in the late amid efforts to rigorously describe light using wave theory, moving beyond the limitations of geometric . In 1882, published his foundational paper "Zur Theorie der Lichtstrahlen," where he applied from to formulate a mathematical solution for diffraction problems in . This approach treated light as satisfying the scalar , enabling the calculation of the diffracted field from boundary values on an . Kirchhoff's work built upon the Helmholtz-Kirchhoff tradition, which extended —originally developed for and hydrodynamics—to acoustic and electromagnetic waves. had earlier formulated the wave equation for propagation in 1860, providing the mathematical foundation for such applications, while Kirchhoff, who joined in as professor of in 1875, adapted these tools to optical . The physical motivation stemmed from the need to explain phenomena, such as the bending of around obstacles, which geometric could not account for; by the 1880s, James Clerk Maxwell's electromagnetic theory of 1865 had established as transverse waves, but a scalar was necessary to simplify for monochromatic waves of wavelength λ, treating the electric field as a scalar ψ satisfying the ∇²ψ + k²ψ = 0, where k = 2π/λ. Despite its influence, Kirchhoff's formulation faced early critiques for mathematical inconsistencies, notably from in 1889—who identified a where the conditions over-specify the problem, potentially leading to discontinuities in the field that violate the —and Lord in the 1890s, who pointed out that the assumed conditions on the and screen did not fully satisfy the everywhere. Rayleigh argued that the theory implied discontinuous field values across the , violating the continuity required by electromagnetic propagation, though it still yielded accurate predictions for many cases. This scalar framework, while approximate, provided a cornerstone for , intuitively extending Huygens' 1678 of secondary wave sources to a quantitative form.

Relation to Huygens' principle

Huygens' principle posits that every point on a serves as a source of secondary spherical wavelets that propagate forward at the speed of the wave, with the new wavefront formed by the to these wavelets. This empirical concept, originally qualitative, provided an intuitive framework for understanding wave propagation and but lacked a rigorous mathematical basis for calculating field amplitudes. Kirchhoff advanced Huygens' idea by transforming it into a quantitative through the application of to the scalar , yielding a boundary integral that computes the diffracted field directly. This approach resolved longstanding issues in Huygens' principle, such as the ambiguous treatment of the obliquity factor, which accounts for the directional dependence of secondary wave contributions, by deriving it systematically from the wave equation rather than ad hoc adjustments. Specifically, Kirchhoff's 1882 incorporated an obliquity factor of \frac{1 + \cos \theta}{2}, enhancing the principle's for phenomena. Under conditions of normal incidence and appropriate approximations, Kirchhoff's diffraction formula aligns with the Huygens-Fresnel principle, incorporating the obliquity factor \frac{1 + \cos \theta}{2}. Physically, this equivalence interprets the diffracted field as arising from secondary wave sources distributed across the aperture surface, with the total field obtained via a that sums their contributions, thereby formalizing Huygens' wavelet superposition in a boundary-value problem context.

Mathematical foundations

Kirchhoff's integral theorem

, a fundamental result in applied to wave propagation, expresses the value of a scalar wave field satisfying the at an interior point in terms of surface integrals over a closed bounding surface. The theorem assumes monochromatic scalar waves of the form u(\mathbf{r}, t) = \Re \left[ U(\mathbf{r}) e^{-i \omega t} \right], where U(\mathbf{r}) satisfies the time-independent \nabla^2 U + k^2 U = 0 with wavenumber k = \omega / c, and c is the wave speed. Additionally, the field must obey the at infinity to ensure outgoing waves, though for the interior problem addressed by the theorem, this supports the boundary behavior. To derive the theorem, consider a volume V bounded by a closed surface S, with observation point P inside V. Let U satisfy the Helmholtz equation in V except possibly at sources, and introduce the Green's function G(\mathbf{r}, \mathbf{r}_P) = \frac{e^{i k R}}{R}, where R = |\mathbf{r} - \mathbf{r}_P| is the distance from P to the field point \mathbf{r} on S; this G also satisfies the Helmholtz equation everywhere except at the singularity \mathbf{r} = \mathbf{r}_P. Apply Green's second identity to U and G over the volume excluding a small sphere \Sigma of radius \epsilon around P: \int_{V - \Sigma} \left( U \nabla^2 G - G \nabla^2 U \right) dV = \oint_{S + \Sigma} \left( U \frac{\partial G}{\partial n} - G \frac{\partial U}{\partial n} \right) dS, where \partial / \partial n denotes the outward normal derivative. Since both functions satisfy the Helmholtz equation in V - \Sigma, the left-hand side vanishes. The integral over S remains, while the contribution from \Sigma evaluates to +4\pi U(P) in the limit \epsilon \to 0, due to the $1/R singularity of G behaving like the fundamental solution of the Laplace equation near P. Thus, rearranging yields the theorem's statement: U(P) = -\frac{1}{4\pi} \oint_S \left[ U \frac{\partial G}{\partial n} - G \frac{\partial U}{\partial n} \right] dS. This formula, first established by Kirchhoff in the context of light rays, holds under the assumptions of scalar wave propagation without absorption and sufficient smoothness of U and the surface S. The theorem provides the mathematical foundation for solving boundary value problems in wave optics by relating interior field values to boundary data.

Boundary value problem in wave optics

In the context of wave optics, the boundary value problem for Kirchhoff's diffraction formula considers an incident plane wave approaching from the source side (z < 0), passing through an aperture in an opaque screen located at the plane z = 0, with the observation point P situated in the diffraction region at z > 0. This geometry models the propagation of scalar light waves through an opening, where the screen blocks transmission except at the aperture, allowing diffracted fields to be evaluated downstream. The Kirchhoff approximation imposes specific boundary conditions to simplify the solution: on the opaque screen, the field satisfies the Dirichlet condition u = 0, while on the , the total field u approximates the incident field u_inc, and its normal derivative ∂u/∂n equals -i k u_inc, where k is the ; this setup neglects back-scattered waves from the screen. These conditions assume the is sufficiently illuminated by the incident wave without significant disturbance from the screen edges. The approximation holds under conditions where the aperture dimensions are much larger than the (ensuring geometric validity locally) but much smaller than the from the aperture to the observation point P (maintaining far-field-like propagation). This introduces a surface of over the aperture alone, as contributions from the screen vanish due to the boundary conditions. The diffracted field at P is then expressed approximately as U(P) \approx \frac{1}{4\pi} \int_{\text{[aperture](/page/Aperture)}} \left[ U \frac{\partial G}{\partial n} - G \frac{\partial U}{\partial n} \right] dS, where G is the for the , and on the , U ≈ U_inc with the corresponding derivative. This integral setup leverages the general Kirchhoff theorem by restricting the integration to the surface.

Derivation of the formula

Application to point source

The application of Kirchhoff's diffraction formula to a establishes the basic framework for calculating the diffracted field in scalar wave . Consider a monochromatic S located at a distance from an opaque screen containing an of arbitrary shape. The source emits a spherical wave that illuminates the , and the observation point P lies in the space beyond the screen. The incident field at any point Q on the surface is given by the spherical wave expression U_{\text{inc}}(Q) = \frac{A e^{i k r_s}}{r_s}, where A is the amplitude, k = 2\pi / \lambda is the wavenumber with wavelength \lambda, and r_s = | \mathbf{r}_S - \mathbf{r}_Q | is the distance from S to Q. To derive the diffracted field at P, Kirchhoff's integral theorem is applied, which stems from Green's second identity for solutions to the Helmholtz equation (\nabla^2 + k^2) U = 0. The theorem expresses the field U(P) in terms of boundary values over a closed surface enclosing P: U(P) = \frac{1}{4\pi} \oint \left( G \frac{\partial U}{\partial n} - U \frac{\partial G}{\partial n} \right) dS, where G is the Green's function G(Q, P) = \frac{e^{i k r_p}}{r_p} representing an outgoing spherical wave from P (with r_p = | \mathbf{r}_P - \mathbf{r}_Q |), and \partial / \partial n denotes the normal derivative outward from the volume. For diffraction, the integration surface is chosen to consist of the aperture A and the screen, but contributions from the opaque screen are assumed zero, leaving only the integral over A. In the Kirchhoff approximation, the and its on the are replaced by those of the undisturbed incident wave, assuming the aperture does not perturb the field significantly. Thus, U(Q) \approx U_{\text{inc}}(Q). The normal of the incident field is approximated as \frac{\partial U_{\text{inc}}}{\partial n} \approx i k \cos \theta_s \, U_{\text{inc}}(Q), where \theta_s is the angle between the incident ray direction at Q (from S to Q) and the surface normal \hat{n} (pointing toward the observation side). Similarly, for the , \frac{\partial G}{\partial n} \approx - i k \cos \theta_p \, G(Q, P), where \theta_p is the angle between the direction from Q to P and \hat{n}. Substituting these into the integral yields U(P) = \frac{i k}{4\pi} \int_A U_{\text{inc}}(Q) \, G(Q, P) \, (\cos \theta_s + \cos \theta_p) \, dS. With U_{\text{inc}}(Q) = \frac{A e^{i k r_s}}{r_s} and G(Q, P) = \frac{e^{i k r_p}}{r_p}, the expression simplifies to the Fresnel-Kirchhoff diffraction formula for a point source: U(P) = \frac{A}{i \lambda} \int_A \frac{e^{i k (r_s + r_p)}}{r_s r_p} \, \frac{\cos \theta_s + \cos \theta_p}{2} \, dS, where the factor arises from \frac{i k}{4\pi} (\cos \theta_s + \cos \theta_p) = \frac{1}{i \lambda} \frac{\cos \theta_s + \cos \theta_p}{2}. The term \frac{\cos \theta_s + \cos \theta_p}{2} is the obliquity (or inclination) factor, which accounts for the directional dependence of wave propagation and ensures no significant backward diffraction; it arises directly from the combination of the normal derivative approximations and represents an average over incident and diffracted directions. This factor is derived by assuming far-field-like behavior for the derivatives while retaining exact distances in the phase terms r_s + r_p, emphasizing the total optical path length in the exponent. For small angles—typical in paraxial approximations where both \theta_s and \theta_p are near zero—the obliquity factor approaches 1, reducing the formula to the Huygens-Fresnel principle: U(P) \approx \frac{A}{i \lambda} \int_A \frac{e^{i k (r_s + r_p)}}{r_s r_p} \, dS. This equivalence highlights how Kirchhoff's approach refines Huygens' wavelet concept by incorporating rigorous boundary conditions from , while the obliquity derivation ensures physical consistency with forward-propagating waves. The full formula with phase terms r_s + r_p captures both near- and far-field effects without further approximation.

Extension to arbitrary sources

The Kirchhoff diffraction formula, originally derived for point sources, can be generalized to an arbitrary incident U_{\text{inc}}(\mathbf{r}_\perp, 0) illuminating an or screen by incorporating the incident field values directly into the boundary integral over the aperture surface \Sigma. This extension leverages the of the wave equation, allowing the diffracted at an observation point \mathbf{r} to be expressed as a superposition of contributions from each point on the aperture, weighted by the local incident field and its . In the scalar approximation, the diffracted field is given by the Fresnel-Kirchhoff integral: U(\mathbf{r}) = \frac{1}{4\pi} \iint_\Sigma \left[ \frac{e^{ikR}}{R} \frac{\partial U_{\text{inc}}(Q)}{\partial n_Q} - U_{\text{inc}}(Q) \frac{\partial}{\partial n_Q} \left( \frac{e^{ikR}}{R} \right) \right] dS_Q, where Q denotes points on the , R = |\mathbf{r} - Q| is the distance from Q to the observation point, k = 2\pi / \lambda is the , and \partial / \partial n_Q is the normal derivative at Q. This form does not assume a specific point-source for the incident wave, making it applicable to complex wavefronts such as Gaussian beams or focused fields. For extended sources, the total diffracted field depends on whether the source emission is coherent or incoherent. In the coherent case, such as illumination of an extended , the total field at the observation point is obtained by integrating the point-source contributions over the source distribution, yielding U_{\text{total}}(P) = \int U_{\text{point}}(P; \xi) \, d\xi, where \xi parameterizes the source points and U_{\text{point}} is the field from an elemental source at \xi. A specific case is uniform illumination, which corresponds to the plane-wave limit where U_{\text{inc}} is constant across the aperture, simplifying the integral to a form proportional to the aperture's transmission function. For incoherent extended sources, such as self-luminous apertures (e.g., thermally emitting objects), the emissions from different source points are uncorrelated, so the total at the point is the incoherent I(P) = \int |U_{\text{point}}(P; \xi)|^2 \, d\xi, rather than the modulus squared of the coherent . Here, the from each elemental source follows I(P) = |U(P)|^2 locally, but global is absent. This approach is common in modeling from rough surfaces or sources. Partially coherent sources, typical of natural light, require handling mutual coherence between source points. The degree of spatial coherence across the aperture is described by the van Cittert-Zernike theorem, which states that for a quasimonochromatic incoherent extended source, the complex degree of coherence \gamma(\mathbf{r}_1, \mathbf{r}_2) between two points is the normalized Fourier transform of the source intensity distribution. In diffraction calculations, this coherence function modulates the Kirchhoff integral, replacing the field U with the mutual intensity J(\mathbf{r}_1, \mathbf{r}_2) = \langle U(\mathbf{r}_1) U^*(\mathbf{r}_2) \rangle, leading to the diffracted mutual intensity as J(P_1, P_2) = \iint J(Q_1, Q_2) K(P_1, Q_1) K^*(P_2, Q_2) \, dQ_1 dQ_2, where K is the Kirchhoff kernel. This framework enables accurate predictions for partially coherent diffraction patterns in optical systems.

Specific diffraction regimes

Fresnel diffraction approximation

The Fresnel diffraction approximation applies to near-field scenarios where the observation point is at a z from the such that the Fresnel number F = a^2 / (\lambda z) \gtrsim 1 (i.e., z \lesssim a^2 / \lambda), with a denoting the size and \lambda the , allowing quadratic phase terms to be retained while neglecting higher-order contributions. This regime describes near-field or intermediate-field , where the wave curvature remains significant, distinguishing it from far-field behavior. Starting from , the derivation involves a binomial expansion of the distance in the of the . The distance from the observation point P at (x, y, z) to an point (x', y', 0) is approximated as |\mathbf{r} - \mathbf{r}'| \approx z + \frac{(x - x')^2 + (y - y')^2}{2z}. For a incident normal to the , the field u \approx 1 (normalized) and its normal derivative \frac{\partial u}{\partial n'} \approx i k u, with the obliquity factor approximated as unity for small angles, and the $1/r amplitude term simplified to $1/z. This yields the quadratic phase expression \exp\left[ i \frac{k}{2z} \left( (x - x')^2 + (y - y')^2 \right) \right], where k = 2\pi / \lambda. The resulting field at P is given by the Fresnel diffraction : U(P) \approx \frac{e^{i k z}}{i \lambda z} \iint U_\text{inc}(x', y') \exp\left[ i \frac{k}{2z} \left( (x - x')^2 + (y - y')^2 \right) \right] \, dx' \, dy', where U_\text{inc}(x', y') is the incident field at the . This integral incorporates the quadratic phase, representing the of the aperture function with a kernel that accounts for curvature. For circular apertures, the Fresnel zone interpretation divides the aperture into concentric regions where the path length difference to the observation point increases by \lambda/2 per zone, each contributing oppositely phased amplitudes that nearly in pairs. The radius of the nth zone is \sqrt{n \lambda b}, with b = z_1 z_2 / (z_1 + z_2) for source-observation distances z_1, z_2; on-axis intensity oscillates with the number of exposed zones, peaking when an odd number is uncovered. A representative example is the single-slit Fresnel pattern, computed via Cornu's spiral, which plots the Fresnel integrals C(v) = \int_0^v \cos(\pi t^2 / 2) \, dt and S(v) = \int_0^v \sin(\pi t^2 / 2) \, dt, with variable v = x \sqrt{2 / (\lambda z)}. For a slit of width a, the features bright "cornua" (horn-like fringes) near the geometric edges due to the spiral's coiled ends, contrasting with the central variation. Another example is the Poisson spot behind a circular opaque disk, where Babinet's principle ensures a bright central maximum in the with matching the unobstructed , arising from constructive of waves diffracting around the obstacle via symmetric Fresnel zones.

Fraunhofer diffraction approximation

The approximation is valid in the far-field regime, where the observation distance z greatly exceeds the characteristic size squared divided by the , i.e., z \gg a^2 / \lambda, and the angular extent of observation \theta is small enough to neglect higher-order terms beyond the in the expansion of the distance. This corresponds to a small Fresnel number F = a^2 / (z \lambda) \ll 1, ensuring that the difference due to terms across the is much less than $2\pi. Under these circumstances, the curvature effects in the are minimal, and the pattern becomes independent of the exact distance z in terms of its angular distribution. The derivation proceeds from the Fresnel-Kirchhoff diffraction by approximating the propagation distance r = \sqrt{z^2 + (x - x')^2 + (y - y')^2} in the \exp(i k r), where k = 2\pi / \lambda. The binomial expansion yields r \approx z + [(x - x')^2 + (y - y')^2] / (2z), but for Fraunhofer conditions, the terms (x^2 + y^2)/(2z) within the are neglected as they contribute negligible phase variation, while the cross terms - (x x' + y y') / z are retained. The obliquity factor is approximated as unity for small \theta. This simplification transforms the integral into a form. The resulting diffracted field at the observation point P(x', y', z) for an incident field U_{\text{inc}}(x, y) over the is \begin{equation} U(P) \approx \frac{e^{i k z}}{i \lambda z} , e^{i k (x'^2 + y'^2)/(2 z)} \iint U_{\text{inc}}(x, y) , e^{-i k (x x' + y y') / z} , dx , dy, \end{equation} where the integral is taken over the . The quadratic outside the integral accounts for the overall spherical curvature at the observation plane but does not affect the angular pattern shape. This expression reveals the Fraunhofer integral as the two-dimensional Fourier transform of the aperture function U_{\text{inc}}(x, y), with spatial frequencies f_x = x' / (\lambda z) and f_y = y' / (\lambda z). The diffraction pattern's intensity is thus | \mathcal{F}\{ U_{\text{inc}} \} |^2, scaled by factors involving \lambda and z, emphasizing the pattern's angular dependence on \theta_x = x'/z and \theta_y = y'/z. In lens-based systems, placing the aperture in the front focal plane of a lens with focal length f performs this Fourier transform directly in the rear focal plane, where the scaling becomes \lambda f instead of \lambda z, enabling practical Fourier optics applications such as spatial filtering. A representative example is the pattern from a with slit spacing d, where the amplitude is the product of the single-slit pattern (a ) and an array factor yielding delta-like peaks at angles \sin \theta_m = m \lambda / d for integer orders m. For a of radius a, the pattern forms the , with normalized intensity I(\theta)/I_0 = \left[ 2 J_1(k a \sin \theta) / (k a \sin \theta) \right]^2, where J_1 is the first-order of the first kind; the central disk contains about 84% of the energy, bounded by the first dark ring at \theta \approx 1.22 \lambda / (2a).

Validity and limitations

Assumptions and boundary conditions

Kirchhoff's diffraction formula relies on the scalar wave approximation, treating the optical field as a scalar quantity that satisfies the , thereby neglecting the vectorial nature of electromagnetic waves and effects. This assumption simplifies the problem to a single scalar but limits accuracy for or vector fields. Additionally, the formula assumes no multiple scattering, considering only single interactions with the or obstacle. The core boundary conditions, known as Kirchhoff's boundary conditions, specify that on the opaque screen, both the field amplitude and its normal vanish, while on the , they match the incident wave exactly. These conditions introduce a discontinuity at the aperture edge, where the field jumps abruptly from zero on the screen to the incident value in the aperture, leading to inherent errors near the boundaries. To ensure physically outgoing , the is imposed, requiring the field to behave like a spherical wave at large distances, decaying as 1/r and satisfying the Sommerfeld condition for no incoming radiation from infinity. The formula is valid when the wavelength λ is much smaller than the aperture dimensions (typically several wavelengths across) and when source and observation distances exceed several wavelengths, ensuring the geometric optics limit holds. However, it fails for grazing incidence at large angles, where errors become significant due to poor handling of near-shadow boundary effects, as seen in comparisons where deviations exceed expectations for incidence angles approaching 90 degrees.

Modern extensions and critiques

One of the primary critiques of Kirchhoff's diffraction formula concerns the inconsistency in its boundary conditions, where the field is prescribed as equal to the incident wave within the and zero outside, while the normal derivative is assumed to be that of the incident wave everywhere on the screen; this leads to discontinuities at the edge that violate the wave equation. emphasized this issue in his analysis of wave passage through apertures, noting that the abrupt change in field values across the boundary creates physical implausibilities in the near-field behavior. In response, and Sommerfeld developed alternative formulations in the late 19th and early 20th centuries that relax one of the boundary conditions to restore consistency, though these still approximate the edge effects. Emil Wolf, in the mid-20th century, addressed edge condition problems by reformulating Kirchhoff's theory to ensure mathematical rigor, demonstrating that the provides an exact solution under specific interpretations of the boundary values for absorbing screens, thereby mitigating the original discontinuities without altering the core . Wolf's work, particularly in collaboration with Marchand during the , showed that the apparent inconsistencies arise from improper application rather than inherent flaws, allowing the theory to align better with experimental observations near edges. Extensions to vector fields have broadened Kirchhoff's scalar framework to handle polarized electromagnetic waves, employing potentials to decompose the electric and magnetic fields into longitudinal and transverse components that satisfy across the diffracting aperture. This vectorial approach, developed in the early and refined through mid-century, accounts for polarization-dependent effects absent in the original scalar theory, enabling accurate predictions for non-paraxial in optical systems. The diffraction wave (BDW) theory represents a significant extension, originating with in 1882 and advanced by Rubinowicz in 1918, who expressed the diffracted field as the superposition of the direct geometrical wave and a cylindrical wave emanating solely from the 's edge contour. Koiter further generalized this in 1950 to three-dimensional geometries, providing a physically intuitive separation of contributions that avoids Kirchhoff's volume integration over the and better handles sharp edges without discontinuities. This formulation has proven particularly useful for high-frequency approximations and aligns with geometrical theories of . In contemporary practice, Kirchhoff's formula is routinely evaluated numerically; for instance, (FFT) algorithms efficiently compute patterns from aperture distributions, while finite-difference time-domain (FDTD) methods incorporate the full Kirchhoff to simulate time-dependent wave propagation in complex environments. These techniques extend the theory's applicability to non-ideal conditions, such as irregular apertures or absorbing media. In antenna theory, the Kirchhoff approximation models radiation from large apertures by treating them as equivalent sources, facilitating pattern predictions for and communication systems. Similarly, in acoustics, it describes by barriers and obstacles, aiding in and design. The validity of solutions derived from Kirchhoff's formula relies on the for the , which requires Sommerfeld's radiation condition at infinity to ensure only outgoing waves and prevent spurious incoming contributions; without this, multiple solutions may exist, particularly in unbounded domains. This condition, introduced in 1912, underscores a limitation in early applications of Kirchhoff's theory that lacked explicit far-field constraints. Refinements to Babinet's principle in the extended its scalar validity—equating patterns from complementary opaque and transparent screens—to vector electromagnetic cases, incorporating and effects to correct discrepancies observed in and optical experiments. These developments, notably in the mid-century works on electromagnetic duality, ensure the principle holds for non-scalar fields under appropriate symmetry conditions.

References

  1. [1]
    Diffraction - Galileo and Einstein
    Scalar Diffraction Theory ... (∇2+k2)ψ(x)=0. Kirchhoff's idea is that we know ψ at S1; it's just that part of the incoming wave that hits the holes (this is a ...
  2. [2]
    None
    Summary of each segment:
  3. [3]
    [PDF] Chapter 15S Fresnel-Kirchhoff diffraction - bingweb
    Jan 22, 2011 · Illustrating the diffraction formula. We assume that the shape of the aperture in the screen is a sphere. Let A be the instantaneous position of ...
  4. [4]
    Zur Theorie der Lichtstrahlen - ADS - Astrophysics Data System
    Export Citation. ADS. Zur Theorie der Lichtstrahlen. Kirchhoff, G. Abstract. Publication: Annalen der Physik. Pub Date: 1882; DOI: 10.1002/andp.18832540409.
  5. [5]
    Zur Theorie der Lichtstrahlen - Kirchhoff - 1883 - Wiley Online Library
    Zur Theorie der Lichtstrahlen. G. Kirchhoff, G. Kirchhoff. Search for more papers by this author. First published: 1883.
  6. [6]
  7. [7]
  8. [8]
    Kirchhoff's theory for optical diffraction, its predecessor and ...
    Aug 10, 2025 · Kirchhoff's 1882 theory of optical diffraction forms the centerpiece in the long-term development of wave optics, one that commenced in the ...
  9. [9]
    27.2 Huygens's Principle: Diffraction – College Physics
    An accurate technique for determining how and where waves propagate is given by Huygens's principle: Every point on a wavefront is a source of wavelets that ...
  10. [10]
    Geometrization of the Huygens–Fresnel principle - AIP Publishing
    May 17, 2024 · 4). In 1882, Kirchhoff demonstrated that a slight modification of Fresnel's diffraction integral to include an obliquity factor was derivable ...
  11. [11]
    [PDF] 7.7 Kirchhoff Theory
    Kirchhoff theory is a rigorous treatment of Huygens' principle, used to compute synthetic seismograms. It is based on Kirchhoff's formula, which computes ...
  12. [12]
    [PDF] 19690014676.pdf - NASA Technical Reports Server
    these are binding on the Kirchhoff Integral Theorem and its vector analogue. ... Born, Max and Emil Wolf, "Principles of Optics", (2nd ed.), Pox anion. Press ...
  13. [13]
    4.1.1 Scalar Diffraction Theory - IuE
    The solution of the diffraction problem is found by specifying Kirchhoff boundary conditions on $ \mathcal {A}$ and $ \mathcal {B}$ , and the Sommerfeld ...
  14. [14]
    [PDF] Diffraction as scattering under the Born approximation - arXiv
    Here, we derive the general formula for the diffracted field generated by an arbitrary incident field at a screen, which includes the aperture above. As a ...
  15. [15]
    None
    Below is a merged summary of the segments from *Introduction to Fourier Optics* by Joseph W. Goodman, consolidating all information from the provided summaries into a comprehensive response. To retain maximum detail, I will use a table in CSV format for each section, followed by a narrative summary where additional context or qualitative information is provided. This approach ensures all equations, page references, and concepts are preserved efficiently.
  16. [16]
    Van Cittert–Zernike Theorem and Lambert’s Law
    **Summary of van Cittert-Zernike Theorem and Diffraction Integrals:**
  17. [17]
    [PDF] Fresnel Diffraction.nb
    In the study of Fresnel diffraction it is convenient to divide the aperture into regions called Fresnel zones. Figure. 1 shows a point source, S, ...Missing: cornua | Show results with:cornua
  18. [18]
    First- and second-order Poisson spots | American Journal of Physics
    Aug 1, 2009 · In this paper we examine the implications of Fresnel's work, particularly the phenomenon of the Poisson spot. The term Poisson spot refers to a ...
  19. [19]
    [PDF] Diffraction
    One plane wave makes an angle of 5 degrees with respect to the z axis, and the second plane wave is along the z-axis. The phase difference between the two given ...Missing: formula | Show results with:formula
  20. [20]
    [PDF] Fourier transforming properties of lenses - MIT OpenCourseWare
    Fourier transforming properties of lenses. • Spherical - plane wave duality. • The telescope (4F system) revisited. – imaging as a cascade of Fourier ...
  21. [21]
    [PDF] Chapter 36 - NJIT
    Diffraction by a Circular Aperture. • The central bright spot in the diffraction pattern of a circular aperture is called the Airy disk. • We can describe ...<|separator|>
  22. [22]
    [PDF] Physical Optics and Diffraction - Princeton University
    Condition for Fraunhofer is that we can neglect the quadratic phase variation with position in the aperture plane. i.e. for all x, or ...
  23. [23]
  24. [24]
    [PDF] Miraculous Success? Inconsistency and Untruth in Kirchhoff's ...
    Marchand and Wolf (1966) show that Kirchhoff's diffraction formula (1) can be derived—exactly—from a consistent set of assumptions. Kirchhoff's boundary ...Missing: original | Show results with:original
  25. [25]
    [PDF] 19730016453.pdf - NASA Technical Reports Server (NTRS)
    At larger angles of incidence, Kirchoff theory fails to represent the scattering process, A shortcoming of Kirchoff theory is that it cannot account for ...
  26. [26]
    Vector diffraction theory for electromagnetic waves
    Traditionally in optics, diffraction is formulated by using only the surface integral (Kirchhoff formulation), where the surface includes the aperture plane. We ...Missing: school | Show results with:school
  27. [27]
  28. [28]
    Generalization of the Maggi-Rubinowicz Theory of the Boundary ...
    Aug 9, 2025 · As a first step towards a generalization of the Maggi-Rubinowicz theory of the boundary diffraction wave, a new vector potential W(Q,P) is ...Missing: Koiter | Show results with:Koiter
  29. [29]
    An FFT‐based Kirchhoff integral technique for the simulation of radio ...
    Mar 6, 2010 · As mentioned in section 4, providing the incident field can be locally regarded as a plane wave, an effective grazing angle can be calculated at ...
  30. [30]
    and far-field calculation in FDTD simulations using Kirchhoff surface ...
    Aug 5, 2025 · Kirchhoff's surface integral representation (KSIR) is used to calculate the near and far fields from the finite difference time domain (FDTD) simulation.
  31. [31]
    Scattering Evaluation of Equivalent Surface Impedances of Acoustic ...
    The magnitudes and directions of such scattered waves can be described by the Helmholtz–Kirchhoff theory of diffraction, which uses Green's theorem to determine ...
  32. [32]
    [PDF] Babinet's Principle for Electromagnetic Fields 1 Problem 2 Solution
    [27] Lord Rayleigh, On the Passage of Waves through Apertures in Plane Screens, and Allied ... [28] Lord Rayleigh, On the Incidence of Aerial and Electric ...
  33. [33]
    Elemental derivation of the Babinet principle in electromagnetic form
    To illustrate the usefulness of the elemental method, we shall describe the diffraction of a normally incident plane wave by a conducting half-plane and tabu-.